insights from inside the volcano: genesis and …

142
The Pennsylvania State University The Graduate School Department of Geosciences INSIGHTS FROM INSIDE THE VOLCANO: GENESIS AND ERUPTION OF THRÍHNÚKAGÍGUR VOLCANICS, REYKJANES PENINSULA, ICELAND A Thesis in Geosciences by Michael R. Hudak 2016 Michael R. Hudak Submitted in Partial Fulfillment of the Requirements for the Degree of Master of Science May 2016

Upload: others

Post on 06-Nov-2021

1 views

Category:

Documents


0 download

TRANSCRIPT

The Pennsylvania State University

The Graduate School

Department of Geosciences

INSIGHTS FROM INSIDE THE VOLCANO: GENESIS AND ERUPTION OF

THRÍHNÚKAGÍGUR VOLCANICS, REYKJANES PENINSULA, ICELAND

A Thesis in

Geosciences

by

Michael R. Hudak

2016 Michael R. Hudak

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Master of Science

May 2016

The thesis of Michael R. Hudak was reviewed and approved* by the following:

Maureen D. Feineman

Assistant Professor of Geosciences

Thesis Co-Advisor

Peter C. LaFemina

Professor of Geosciences

Thesis Co-Advisor

Tanya Furman

Professor of Geosciences

Associate Vice President and Associate Dean for Undergraduate Education

Michael Arthur

Professor of Geosciences

Interim Head of Graduate Programs of the Geosciences Department

*Signatures are on file in the Graduate School

iii

ABSTRACT

Thríhnúkagígur is the main cone of a small 3500 ka fissure eruption in the

Brennisteinsfjöll fissure swarm on the Reykjanes Peninsula. Beneath the summit of the main cone

is a 120 m deep cave exposing a large part of the uppermost magmatic plumbing system and a

buried tephra cone. The exposure of this structure offers a unique view into a non-unique eruptive

process. Basaltic fissure eruptions are common in many tectonic and volcanic settings. Direct

observations in the cave in addition to major element, trace element and Sr-Nd-Pb isotope

analyses allow a fissure eruption on the Reykjanes Peninsula to be investigated from melt

generation in the mantle to eruption. Field relationships between the main dike that fed the

eruption and the unconsolidated buried tephra suggest that the tephra was entrained by the

erupting magma. Assuming the cave represents the volume of tephra entrained and removed from

the system, estimates of the erupted volume of lava produce maximum estimates of ~29% tephra

content in the lavas. Furthermore, the location of the buried tephra directly beneath the main cone

of Thríhnúkagígur suggest that the buried tephra acted as a local asperity in the crust forming a

preferential pathway for the magma to ascend to the surface. The main vents were developed here

because flow was focused by the buried tephra. However, the vertical sheet-like structure of the

dike indicates that the far field stresses dictate the orientation of fissure despite the contrast in

competency between the buried tephra and overlying basaltic lava flow.

The Sr-Nd-Pb isotope systematics illustrate that Thríhnúkagígur lavas are consistent with

other relatively enriched compositions observed on the Reykjanes Peninsula, with radiogenic Pb

and Sr, and relatively unradiogenic Nd. These isotopes suggest that Thríhnúkagígur magmas are

derived from two main components – an enriched plume source and a depleted mantle source –

with possible minor contributions from an EM1-like mantle source. Anomalously high ratios of

Nb/U at Thríhnúkagígur suggest that despite isotopic similarities between the Reykjanes

Peninsula and the Eastern Volcanic Zone (EVZ), the enriched plume source at the Reykjanes is

unique from the EVZ in its enriched Nb. Decoupling of the Sr isotopes from Pb and Nd isotopes

suggests that the source of radiogenic Sr at Thríhnúkagígur may be assimilation of basaltic crust

that has been hydrothermally altered by seawater. The tephra entrainment hypothesis is tested

with immobile trace element mixing models. Models of La/Yb vs. Nb/Zr yield maximum

estimate 28-35% of the tephra component in the lavas, consistent with results from the

geophysical analysis. However, large uncertainty in the initial magma composition limits the

precision of this model and some samples may not contain any entrained tephra. Finally, the

variability in the major and trace element geochemistry and the Sr isotopes, particularly in the

uppermost magmatic plumbing system, suggest that the small volume of erupted magma was not

homogenized prior to eruption. In the absence of a central volcano at Brennisteinsfjöll, we invoke

a network of dikes and sills that are quasi-isolated until the eruption, allowing for variable

geochemical evolution. This model both preserves magma heterogeneity from the mantle and

facilitates the evolution of magmas with variable compositions, consistent with the variation in

geochemistry observed at multiple levels in this single eruption.

iv

TABLE OF CONTENTS

List of Figures .......................................................................................................................... vi

List of Tables ........................................................................................................................... viii

Acknowledgements .................................................................................................................. ix

Chapter 1 Geologic Overview ................................................................................................. 1

1.1 Introduction ................................................................................................................ 1 1.2 Regional tectonic setting ............................................................................................ 3

1.2.1 The Mid-Atlantic Ridge and the Icelandic plume ........................................... 3 1.2.2 Neovolcanic zones ........................................................................................... 4 1.2.3 Deglaciation and volcanism ............................................................................ 6

1.3 Regional geochemical overview ................................................................................ 7 1.4. The Reykjanes Peninsula and the Brennisteinsfjöll volcanic system ........................ 11

1.4.1 Reykjanes Peninsula tectonics and volcanism ................................................ 11 1.4.2 Reykjanes Peninsula lava chemistry ............................................................... 13

References ........................................................................................................................ 21

Chapter 2 Inside the Volcano: Rapid entrainment of unconsolidated crust during a

monogenetic fissure eruption ........................................................................................... 26

2.1 Introduction ................................................................................................................ 26 2.2 Field descriptions ....................................................................................................... 27

2.2.1 Surface geology ............................................................................................... 27 2.2.2 Subsurface Geology ........................................................................................ 29

2.3 Methods ...................................................................................................................... 34 2.3.1 Light Detection and Ranging (LiDAR) ........................................................... 34 2.3.2 Density Measurements .................................................................................... 35

2.4 Results ........................................................................................................................ 35 2.5 Discussion .................................................................................................................. 36

2.5.1 Cave formation ................................................................................................ 36 2.6 Conclusions ................................................................................................................ 39 References ........................................................................................................................ 58

Chapter 3 Magma genesis at Thríhnúkagígur, Reykjanes Peninsula, Iceland: Implications

for magmatic plumbing structure at monogenetic volcanic fields ................................... 60

3.1 Introduction ................................................................................................................ 60 3.2 Sample descriptions ................................................................................................... 61

3.2.1 Surface lavas ................................................................................................... 62 3.2.2 Dikes ............................................................................................................... 62 3.2.3 Plugged conduit ............................................................................................... 63 3.2.4 Buried tephras ................................................................................................. 63

3.3 Methods ...................................................................................................................... 64

v

3.3.1 Point counting ................................................................................................. 64 3.3.2 Major element analytical methods .................................................................. 65 3.3.3 Trace element analytical methods ................................................................... 65 3.3.4 Pb, Sr, and Nd isotope analytical methods ...................................................... 67

3.4 Results ........................................................................................................................ 68 3.4.1 Major elements ................................................................................................ 68 3.4.2 Trace elements................................................................................................. 69 3.4.3 Pb, Sr, and Nd isotopes ................................................................................... 71

3.5 Discussion .................................................................................................................. 72 3.5.1 Source geochemistry ....................................................................................... 72 3.4.2 Magmatic processes ........................................................................................ 82 3.5.3 Brennisteinsfjöll plumbing system .................................................................. 89

3.6 Conclusions ................................................................................................................ 92 References ........................................................................................................................ 123 Appendix: Isotope geochemistry procedures ................................................................... 128

vi

LIST OF FIGURES

Figure 1.1. Map of Iceland and its neovolcanic zones. These zone trace the plate

boundary between North America (NA) and Eurasia (EU). KR = Kolbeinsey Ridge;

TFZ = Tjörnes Fracture Zone; NVZ = Northern Volcanic Zone; EVZ = Eastern

Volcanic Zone; MIB = Mid-Iceland Belt; WVZ = Western Volcanic Zone; RP =

Reykjanes Peninsula; RR = Reykjanes Ridge. Volcanic zones are composed of many

individual volcanic systems (Einarsson and Sæmundsson, 1987). Contours mark the

approximate depth to the Moho as determined by Darbyshire et al. (2000) and the

map is modified from Brounce et al. (2012). ................................................................... 16

Figure 1.2. Variations along the Mid-Atlantic Ridge in La/Sm and Nb/Zr. Values are <1

for La/Sm on the South Reykjanes Ridge (SRR) and the Kolbeinsey Ridge (KR) and

<0.05 for Nb/Zr. These ratios are greater than 1.5 and 0.08 for mainland Iceland (I).

The North Reykjanes Ridge (NRR) is transitional between ambient MORB

compositions and those in lavas derived from the Icelandic plume. Data from

(Hanan et al., 2000).. ........................................................................................................ 17

Figure 1.3. Map of the Reykjanes Peninsula showing the four fissure systems (shaded

grey). From east to west they are the Reykjanes, Krísuvík, Brennisteinsfjöll, and

Hengill volcanic systems. Thríhnúkagígur is located in Brennisteinsfjöll. Reykjavík

and Thríhnúkagígur are shown in black stars. Outlines of the volcanic systems are

from Einarsson & Sæmundsson (1987). .......................................................................... 18

Figure 1.4. Enriched historic Reykjanes Peninsula lavas plotted on Nb/Zr vs. 87Sr/86Sr

(a) and Nb/Zr vs. 143Nd/144Nd. Data from Peate et al. (2009) fall into “more

enriched” and “less enriched” categories. More enriched sample correlate with lavas

from Brennisteinsfjöll, while less enriched data are from the eastern Reykjanes

Peninsula fissure swarms of Reykjanes and Krísuvík (b). ............................................... 19

Figure 1.5. Changes in lava chemistry on the Reykjanes Peninsula through time from Gee

et al. (1998b). Plots of lava age vs. Mg# (a), 87Sr/86Sr (b), Nb/Zr (c), and 143Nd/144Nd

(d). The period during and immediately following deglaciation (13-9 ka) produced

more primitive, depleted magmas with higher Mg# and 87Sr/86Sr and lower Nb/Zr

ratios and 143Nd/144Nd....................................................................................................... 20

Figure 2.1. Map of the Thríhnukar fissure system modified from Sæmundsson (2006).

Samples are shown in black stars. Subterranean samples are excluded from this map,

but are generally located underneath the Thríhnúkagígur spatter cone in red. ................ 43

Figure 2.2. Two vents comprise the main cone of Thríhnúkagígur. The northern vent

(left) became plugged towards the end of the eruption, while the southern vent

(right, with scaffolding) remained open, providing an entrance to the cave below.

Photo from P. LaFemina. ................................................................................................. 44

Figure 2.3. Succession of thin (2-5 cm) lava flows back into the vent at the summit of

Thríhnúkagígur. Photo from P. LaFemina. ...................................................................... 45

vii

Figure 2.4. LiDAR scan of the interior of the cave at Thríhnúkagígur. Green and blue

coloring represents different reflectivities of the rock. The dike, plugged conduit,

and the buried tephra appear in blue. The image on the left is a view parallel to the

orientation of the dike. The view on the right is perpendicular to the orientation of

the dike. ............................................................................................................................ 46

Figure 2.5. Magma flowback rivulets on a fallen block in the cave. Photo from P.

LaFemina. ........................................................................................................................ 47

Figure 2.6. Stalactites of lava that dripped back into the cave during and/or following the

eruption. Photo from P. LaFemina. .................................................................................. 48

Figure 2.7. The NNE cave wall. A photo mosaic by P. LaFemina is simplified in a

cartoon depicting the dike and tephra with sample locations. .......................................... 49

Figure 2.8. Representative buried tephra (right; TNG-14-15) is poorly consolidated

compared to the well-welded hyaloclastite (left) collected at the from the gentle

dipping hyaloclastite unit postulated by Sæmundsson (2006). ........................................ 50

Figure 2.9. A panorama of the dike at the north end of the cave. Here it cross-cuts the

tephra apron and to the left interfingering between the buried tephra and subsequent

lava flows can be seen. Photo from P. LaFemina. ........................................................... 51

Figure 2.10. Photograph of the interior of the cave. The plugged conduit is the roundish

purple feature on the left. The dike is the vertical purple-black feature on the right.

The buried tephra cone on the bottom left if black. ......................................................... 52

Figure 2.11. A ~0.5 m long gas pocket in the center of the dike. ............................................ 53

Figure 2.12. LiDAR scan of the Thríhnúkagígur cone and cave with the location of the

largest of the proto-conduits indicated (a). A LiDAR scan of the main proto-conduit

shows its geometry in greater detail (b). .......................................................................... 54

Figure 2.13. Stages of Thríhnúkagígur formation starting with the formation (a) and

burial of the tephra cone (b), followed by dike injection (c) and eruption along a

fissure (d). As the eruption continues, the magma both progressively entrains more

tephra from depth and narrows from a fissure into a couple of vents (e anf f). Part g

shows the present day outline of the cave. The left hand side of the diagram shows a

cross-section within the plane of the dike. The left shows a cross-section

perpendicular to the dike. ................................................................................................. 55

Figure 3.1. Photomicrographs of two phenocryst population, a more tabular euhedral one

(a) and a high aspect ratio one (b) in lava TNG-14-14m; the phenocryst-poor chilled

margin (c; TNG-14-27e) and flow banding with interstitial clinopyroxene in the

interior the same dike (d; TNG-14-27i); aphanitic, phenocryst-poor plugged conduit,

TNG-14-18 (e); and large olivine phenocrysts from the Miðhnúkar lava flow, TNG-

14-35 (f). .......................................................................................................................... 106

viii

Figure 3.2. Total alkali-silica plot (TAS). All Thríhnúkagígur samples analyzed are

subalkaline basalts. Buried tephra samples are represented by triangle; lava and dike

samples from the Holocene eruption are circles; the lava sample from Miðhnúkar is

a cross. TAS diagram from Le Bas and Le Maitre (1986). .............................................. 107

Figure 3.3. Fenner diagrams of MgO vs. the major oxides and chemical index of

alteration (CIA). Circles represent lavas and dikes. Triangles represent buried tephra

samples. The cross is the Miðhnúkar lava. Open symbols indicate samples with a

high CIA samples likely to be affected by secondary alteration. ..................................... 108

Figure 3.4. Plots of MgO vs. trace elements and CaO/Al2O3. Circles represent lavas and

dikes while triangles represent buried tephra samples. The cross is the lava from

Miðhnúkar. Open symbols indicate samples with a high CIA and likely to be

affected by secondary alteration. ...................................................................................... 109

Figure 3.5. Primitive mantle normalized multielement diagram for the main sample suites

and unique samples from Thríhnúkagígur. All samples are characterized by small

positive Ba and U, and relatively large Nb and Ta anomalies. All samples have

moderate to large negative Pb anomalies, except for the vent spatter (MC-612-03),

which is thought to be a result of contamination. Most other trace elements display

similar patterns across regardless of sample type, except for Sr which has a positive

anomaly in relatively more depleted plugged conduit (TNG-14-18) and chilled

margin of the dike (TNG-14-27e) and a negative anomaly in the more enriched

buried tephra. Primitive mantle normalization values are from Sun and McDonough

(1989). .............................................................................................................................. 110

Figure 3.6. Chondrite normalized rare earth element (REE) spider diagram by rock type.

All samples have a gently sloping negative trend with higher LREE values than

HREE values. The chilled margin of the dike (TNG-14-27e) has a weak positive Eu

anomaly. Chondrite normalization values are from Sun and McDonough (1989). ......... 111

Figure 3.7. Plots of 206Pb/204Pb vs. 208Pb/204Pb (a) and 207Pb/204Pb (b). Circles represent

lavas and dikes while triangles represent buried tephra samples. The cross is the lava

from Miðhnúkar. Open symbols indicate samples with a high CIA. Error bars

represent either the in run instrumental error or the standard reproducibility,

whichever source of error is larger.. ................................................................................. 112

Figure 3.8. Plot of 87Sr/86Sr vs. 143Nd/144Nd. The buried tephras have consistently lower 143Nd/144Nd ratios compared to the lavas and dikes across the same range of 87Sr/86Sr

ratios. Three dike samples have the least radiogenic Sr isotope ratios at the same

relative radiogenic Nd isotope ratio. Circles represent lavas and dikes while triangles

represent buried tephra samples. The cross is the lava from Miðhnúkar. Open

symbols indicate samples with a high CIA. Error bars represent either the in run

instrumental error or the standard reproducibility, whichever source of error is

larger. ............................................................................................................................... 113

Figure 3.9. (La/Yb)N vs. Nb/Zr. Reykjanes Peninsula and the Western Volcanic Zone

lavas in SW Iceland extend to similarly high Nb/Zr ratios as lavas in the Eastern

Volcanic Zone, but at much lower (La/Yb)N ratios. Thríhnúkagígur lavas and dikes

ix

are consistent other lavas from SW Iceland while the buried tephra record some of

the most enriched values in the region. Circles represent lavas and dikes while

triangles represent buried tephra samples. The cross is the lava from Miðhnúkar.

Open symbols indicate samples with a high CIA. Data for Vestmannaeyjar are from

Furman et al. (1991), Kokfelt et al. (2006), and Peate et al. (2010); EVZ data are

from Kokfelt et al. (2006), Peate et al. (2010), and Manning and Thirlwall (2014);

WVZ data are from Kokfelt et al. (2006) and Koornneef et al. (2012); and data from

the Reykjanes Peninsula are from Kokfelt et al. (2006), Peate et al. (2009), and

Koornneef et al. (2012). ................................................................................................... 114

Figure 3.10. Plots of Nb vs. Nb/U (a) and Ce vs. Ce/Pb (b). Global ocean island basalts

and mid-ocean ridge basalts yield Nb/U ratios of 47 ± 10 at all concentrations of Nb

and Ce/Pb ratios of 25 ± 5 at all Ce concentrations (Hofmann et al., 1986). Dashed

lines show these ranges. Lavas in SW Iceland on the Reykjanes Peninsula and in the

WVZ yield much higher Nb/U and Ce/Pb than observed elsewhere, not just in

Iceland, but globally. Circles represent lavas and dikes while triangles represent

buried tephra samples. The cross is the lava from Miðhnúkar. Open symbols indicate

samples with a high CIA. Data for Vestmannaeyjar are from Kokfelt et al. (2006)

and Peate et al. (2010); EVZ data are from Sigmarsson et al. (1992), Kokfelt et al.

(2006), Peate et al. (2010), and Manning and Thirlwall (2014); WVZ and RP data

are from Kokfelt et al. (2006) and Koornneef et al. (2012). ............................................ 115

Figure 3.11. Plots of 206Pb/204Pb vs. 208Pb/204Pb with proposed mantle end members from

Thirlwall et al. (2004) and Kitagawa et al. (2008). South Iceland lavas form an array

primarily between an enriched and a depleted end member on both plots.

Thríhnúkagígur samples are broadly consistent with composition observed on the

Reykjanes Peninsula and Brennisteinsfjöll. They are pulled slightly off the NHRL

(Hart, 1984) towards an EM1-like end member. Data for the southern Reykjanes

Ridge (SRR) are from Thirlwall et al., 2004; Reykjanes, Krísuvík, and

Brennisteinsfjöll data are from Thirlwall et al., 2004, Kokfelt et al., 2006, and Peate

et al., 2009; Hekla and Katla data are from Park, 1990 and Furman et al., 1995;

Vestmannaeyjar data are from Furman et al., 1991, Chauvel and Hémond, 2000,

Kokfelt et al., 2006, and Peate et al., 2010. ...................................................................... 116

Figure 3.12. Nb/U plotted against 206Pb/204Pb (a), 208Pb/204Pb (b). Solid black triangles

represent Thríhnúkagígur buried tephra. Solid black circles represent

Thríhnúkagígur lavas and dikes. Grey circles represent other Reykjanes Peninsula

lavas from Kokfelt et al. (2006). ...................................................................................... 117

Figure 3.13. A plot of 87Sr/86Sr vs. 143Nd/144Nd. Thríhnúkagígur samples are more

consistent with the Reykjanes fissure swarm and Vestmannaeyjar than

Brennisteinsfjöll and are influenced by an EM1-like component, or more likely,

HAC. Sr-Nd isotopes, data for the southern Reykjanes Ridge (SRR) are from

Thirlwall et al., 2004; Reykjanes, Krísuvík, and Brennisteinsfjöll data are from

Thirlwall et al., 2004, Kokfelt et al., 2006, and Peate et al., 2009; Hekla data are

from Park, 1990, Sigmarsson et al., 1992, and Furman et al., 1995; Katla data are

from Park, 1990 and Furman et al., 1995; Vestmannaeyjar data are from Furman et

al., 1991 and Kokfelt et al., 2006. .................................................................................... 118

x

Figure 3.14. A plot of 206Pb/204Pb vs. 87Sr/86Sr. Thríhnúkagígur samples are most

consistent with the lavas with highly radiogenic Sr isotopes from the Reykjanes

fissure swarm. Data for the southern Reykjanes Ridge (SRR) are from Thirlwall et

al., 2004; Reykjanes, Krísuvík, and Brennisteinsfjöll data are from Thirlwall et al.,

2004, Kokfelt et al., 2006, and Peate et al., 2009; Hekla data are from Park, 1990,

Sigmarsson et al., 1992, and Furman et al., 1995; Katla data are from Park, 1990 and

Furman et al., 1995; Vestmannaeyjar data are from Furman et al., 1991 and Kokfelt

et al., 2006. ....................................................................................................................... 120

Figure 3.15. Sr-Nd isotope mixing model for Thríhnúkagígur lavas and dikes. First, the

dike sample with the least radiogenic Sr isotope composition (TNG-14-24, purple

circle) is mixed with a strongly radiogenic hydrothermally altered crustal end

member from Gee et al., 1998 (blue cross). Then mixing occurs between various

points along the first curve and the least radiogenic buried tephra sample, marked

with a red triangle. ........................................................................................................... 121

Figure 3.16. Incompatible trace element ratio mixing curve, Nb/Zr vs. La/Yb. The

average buried tephra (red diamond) of the representative samples forms the

enriched end member. The geochemistry of the intrusive part of the system is

heterogeneous, so three possible end members are used (yellow diamonds): the

plugged conduit (TNG-14-18), the chilled margin of the dike (TNG-14-27e), and

isotopically most depleted dike sample (TNG-14-24) used as an end member in

Figure 3.13. For the latter two primitive end members, green crosses on the mixing

curves represent 20% buried tephra, blue crosses represent 10%, and red crosses

represent 5%. Dikes and lavas representing possible mixtures of the buried tephra

and original magma composition are blue and yellow circles. ........................................ 122

Figure 3.17. Cartoon of the magmatic plumbing system below Brennisteinsfjöll. The

depth to the base of the crust is to scale, but the volcanic features, dikes, and sills are

not to scale. Hydrothermally altered crust or hyaloclastite layers (HC) likely play a

role in slowing magma ascent. Orange stippled zones represent sills or sill-like

intrusions that have stalled at low density layers. ............................................................ 123

xi

LIST OF TABLES

Table 2.1. Estimates of Thríhnúkagígur lavas volumes and calculation of tephra contents. ... 42

Table 2.2. Densities of Thríhnúkagígur lava and buried tephras. ............................................ 42

Table 3.1. Mineral and groundmass modes in percent. Modes are normalized to 100%

after removing the vesicle mode, which is reported separately. ...................................... 96

Table 3.2. Major element chemical analyses in weight percent oxides. Data were

collected by a Perkin-Elmer Optima 5300DV ICP-AES for a representative sample

set for the units present at Thríhnúkagígur. Mg# and the chemical index of alteration

(CIA) are unitless numbers. Mg# is defined as 100 * Mg/(Mg + Fe) where Mg and

Fe are molar abundances. CIA is defined as 100 * Al2O3/(Al2O3 + CaO + Na2O +

K2O). For plotting major element compositions, all samples were normalized to

100% after removing LOI. *Duplicate analysis. Duplicates are averaged for all

figures............................................................................................................................... 97

Table 3.3. CIPW norms. Calculated from major element data and trace element

concentrations of Sr, Ba, Cr, Ni, and Zr. Samples TNG-14-22, TNG-14-28, and MC-

612-03 are excluded from this table because they have been altered to an extent that

makes their bulk chemistry fall outside the range of compositions for which norms

can be calculated. ............................................................................................................. 99

Table 3.4. Trace element chemical analyses. Data were collected by a Thermo X-Series II

Quadruple ICP-MS for samples from the Thríhnúkar fissure system. All

concentrations are in ppm. NR = not reported. *Duplicate analysis. ............................... 100

Table 3.5. Pb-Sr-Nd isotope ratios. Data were collected by Finnigan MAT 262 multi-

collector thermal ionization mass spectrometer (TIMS). Two standard deviations are

given in ten-thousandths for Pb ratios and one-millionths for 87Sr/86Sr and 143Nd/144Nd. ...................................................................................................................... 103

Table 3.6. Mixing end member compositions for Figures 3.14 and 3.15. HAC =

hydrothermally altered crust; from Gee et al. (1998a). .................................................... 105

xii

ACKNOWLEDGEMENTS

First of all, this thesis would not have been possible without generous funding from the

Department of Geosciences. Thank you to Hiroshi and Koya Ohmoto, the family of Charles E.

Knopf, and the Hess Corporation for graciously endowing the department thereby allowing

students like myself to conduct research in the field. And thanks to the NSF for funding my

isotopic analyses through an international collaboration grant.

I must also thank Björn Òlafsson and all of the folks at Inside the Volcano for granting

me access to the cave below Thríhnúkagígur. The title of this thesis is in honor of your assistance.

Thanks also to Jim Normandeau at UNAVCO for providing me with such an excellent

LiDAR data set.

To Halldór, your assistance in the field was invaluable. I also deeply appreciate your

input throughout this project and your willingness to share your experience with me.

Samuele, thanks for teaching me ion chromatography and TIMS methodology. I learned a

ton from you. Grazie mille also for being so hospitable. You were a fantastic host!

To my parents, Mark and Bonnie, you’re the greatest. Thanks for all your emotional (and

occasional financial) support of my endeavors. I cannot describe how much it means to me.

These last three years have been truly enjoyable largely because of the great people I

have been able to call my friends. It was not always a smooth ride, but your incredible support

helped me with the road bumps along the way. Thank you guys! Nick and Helen, I especially do

not know what I would have done without you.

Tanya, I appreciate the tough love. I really believe it has forced me to address some of

my weakness and helped me to become a better scientist. Thanks for being a skeptic and asking

me tough questions.

Pete, thanks for forcing me to step outside of my comfort zone. You have helped me to

become more comfortable incorporating geophysics into my research. Thanks also for allowing

me to take on this project! It was a great opportunity and an exciting for me.

And to Maureen, thanks for being so accessible and helpful. I am certain that I have tried

your patience during this process, so thanks for putting up with me! You’ve given me some great

opportunities and connections, for which I cannot express my appreciation enough. Thank you.

Chapter 1

Geologic Overview

1.1 Introduction

The goal of this project is to better understand the origin, evolution, and

ultimately the eruption of basalts on the Reykjanes Peninsula in Iceland in the absence of

central volcanoes or any large homogenizing magma chambers. Direct observations of

the shallow magmatic plumbing system in a cave below the main vent of the fissure

eruption at Thríhnúkagígur, together with surficial structures form the basis for

deciphering eruptive style and process. Whole rock chemical and isotopic compositions

of erupted lavas, their corresponding intrusive units, and some of the intruded country

rock provide insights into the mantle sources for Thríhnúkagígur magmas and the crustal

processes that drove their evolution.

Chapter 1 is an overview of the geodynamics and volcanism in Iceland that

provides context for the following chapters. With no eruptions on the Reykjanes

Peninsula since the 13th century, the best analogs for eruptive process are other fissure

eruptions in Iceland. Yet, these comparisons are imperfect because the Reykjanes

Peninsula lacks central volcanoes or large magma storage chambers. In addition, few

eroded cones or vents exist on the Reykjanes Peninsula because it has been largely

resurfaced by Holocene lava flows since the end of the last glaciation. Few places,

therefore, expose the shallow plumbing system of fissure eruptions that would reveal

2

information about eruptive processes and vent development. Thríhnúkagígur, the main

vent of a monogenetic fissure eruption in the Brennisteinsfjöll fissure swarm, is a unique

place to study eruptive process on the Reykjanes Peninsula. The uppermost ~120 meter

of the conduit remained evacuated following the eruption, revealing a cross-section

though the shallow magmatic plumbing system.

In Chapter 2, direct observations of the internal structure of the Thríhnúkagígur

magmatic plumbing system and surface geology forms the basis for describing the

eruption, the development of the vent, and the formation of the cave. While one goal of

this project is to understand what processes might be common for Reykjanes Peninsula

fissure eruptions, the lung-shaped cave below the main vent is highly unusual. This

structure warrants consideration of how the eruption at Thríhnúkagígur is also

uncharacteristic of monogenetic fissure eruptions in this setting. A buried tephra cone is

preserved at depth in the cave. The potential incorporation of this unconsolidated material

into the erupting magma is investigated in this chapter.

With constraints on eruptive process in the uppermost crust, geochemical and

isotopic analyses in Chapter 3 are used to evaluate the mantle sources of these magmas as

well as their storage and evolution in the crust between where melts were generated and

where they were erupted. Lavas from the Reykjanes Peninsula have been well-

characterized geochemically, but seldom have individual eruptions with both intrusive

and extrusive samples been studied in detail. A comprehensive whole rock major and

trace element and Sr-Nd-Pb isotope data set is herein presented. Competing views of

unadulterated mantle melts and significant contamination of these melts in crustal magma

3

chambers have been presented for Reykjanes Peninsula lavas in previous literature.

Chapter 3 evaluates these ideas by investigating a single small volume eruption in detail.

1.2 Regional tectonic setting

1.2.1 The Mid-Atlantic Ridge and the Icelandic plume

Iceland is a remarkably active volcanic province that straddles the Mid-Atlantic

Ridge (MAR). This substantially thickened, buoyant oceanic crust is the result of

increased melt production related to the concurrence of a mantle plume with the MAR.

Its location at a divergent plate boundary has caused the Icelandic hotspot track to trail

both east to the Faroe Islands on the Eurasian plate and west to Greenland on the North

American plate. These tracks suggest that the Icelandic hotspot is at least as old as the

opening of the North Atlantic around 56 Ma, as determined by magnetic anomalies

(Nunns et al., 1983) and the geochronology of flood basalts and picrites in eastern

Greenland (Larsen et al., 1992; Saunders et al., 1997).

Numerous geophysical studies, such as ICEMELT (Bjarnason et al., 1996; Wolfe

et al., 1997) and HOTSPOT (Foulger et al., 2000), have used comprehensive seismic

tomography surveys to image the plume beneath Iceland. These studies and others

converge on a vertical, cylindrical thermal anomaly of approximately 150° C above

ambient mantle temperature extending to at least 400 km depth in the mantle (Bjarnason

et al., 1996; Wolfe et al., 1997; Foulger et al., 2000; Bjarnason, 2008; Brandsdóttir and

Menke, 2008). The lack of an observable, deep-seated mantle plume extending to the 660

km discontinuity or D” beneath Iceland has led some authors to create alternative models

4

for enhanced melt production in Iceland. Foulger and Anderson (2005) argue that the

seismically observed structure can be explained by a complex tectonic model, whereby

subducted oceanic crust from the closure of the Iapetus Ocean is trapped in Laurasian

continental lithosphere, giving rise to an especially fertile mantle in a location where the

geotherm is already elevated due to the presence of the MAR. However, the vast majority

of data for Iceland have been interpreted within a hotspot framework since the advent of

the fixed mantle hotspot hypothesis by Wilson (1963) and Morgan (1971). The work

presented herein likewise assumes the existence of a mantle plume with geochemical

properties similar to those of an ocean island basalt (OIB) source when discussing mantle

sources.

The locus of the plume is assumed to be under the thickest crust, which reaches a

maximum of approximately 40 km thick beneath Europe’s largest ice cap, Vatnajökull,

slightly to the east of Iceland’s center (Figure 1.1). The average crustal thickness is 25-26

km (Darbyshire et al., 1998; Darbyshire et al., 2000; Björnsson et al., 2005; Bjarnason

and Schmeling, 2009), whereas the thinnest crust is 18-20 km thick (Figure 1.1) and may

be as thin as 14 km where the Reykjanes Peninsula goes offshore (Weir et al., 2001).

1.2.2 Neovolcanic zones

Much of the present day magmatism in Iceland exists within four main

neovolcanic zones. Three of these are rift zones that comprise the subaerial MAR: the

Western Volcanic Zone (WVZ), the Eastern Volcanic Zone (EVZ), and the Northern

Volcanic Zone (NVZ). The fourth major volcanic zone, the Reykjanes Peninsula (RP), is

5

a transtensional plate boundary that connects the southern tip of the WVZ to the

submarine Reykjanes Ridge (Figure 1.1).

The plate boundaries and volcanic zones shown in Figure 1.1 are herein described

from north to south. In the modern plate configuration, the submarine Kolbeinsey Ridge

north of Iceland connects to the NVZ by the Tjörnes Fracture Zone. The NVZ ends

beneath Vatnajökull and two rift zones continue to the southwest. The EVZ forms a

continuous boundary with the NVZ that extends south of Vatnajökull. The WVZ is offset

from the NVZ to the west by the E—W oriented Mid-Iceland Belt extending from

Vatnajökull to Langjökull. From there, the WVZ extends to the southwest where it

terminates at the eastern end of the transtensional Reykjanes Peninsula. The RP consists

of a series of NE-SW trending parallel volcanic systems aligned in an E—W lineament

connecting the subaerial expression of the MAR at the WVZ to the submarine Reykjanes

Ridge to the west. The EVZ and the WVZ are connected in the south by a young non-

transform fault system, the South Iceland Seismic Zone (SISZ). The SISZ extends east

from the Hengill volcanic system on the Reykjanes Peninsula and intersects the EVZ

between Hekla volcano and Torfajökull caldera (Geirsson et al., 2012). Additionally,

there are two off-rift volcanic centers, Öræfajökull in the east and Snæfellsnes in the

west.

Spreading of Iceland’s rift zones occurs at a rate of ~17-18 mm/yr total in the

MORVEL model (DeMets et al., 2010), making it an ultraslow spreading center.

LaFemina et al. (2005) used geodetic GPS measurements to characterize spreading rates

along axis of the WVZ and EVZ. They determined that rates in the EVZ decrease to the

south along the rift zone from the full North American-Eurasian spreading rate at the

6

juncture with the NVZ (19.0 ± 2.0 mm/yr) to 11.0 ± 0.8 mm/yr in the southern EVZ. This

decrease correlates with a spreading rate increase moving south along the WVZ (2.6 ±

0.9 mm/yr to 7.0 ± 0.4 mm/yr). The EVZ now accommodates two-thirds of the MAR

spreading rate in south Iceland (LaFemina et al., 2005), which has led some authors to

suggest that the WVZ is dying out as another ridge jump is made from the WVZ to the

EVZ (Hardarson et al., 1997). The North American-Eurasian plate boundary continues to

reconfigure to the southeast as the plates drift WNW over the stationary Icelandic plume

(Hardarson et al., 1997; Mittelstaedt et al., 2008; Walters et al., 2013). In this framework,

the GPS geodetic data of LaFemina et al. (2005) demonstrate that ridge reconfiguration is

a transitional process rather than an instantaneous one as the EVZ continues to propagate

to the south in the present day.

1.2.3 Deglaciation and volcanism

The Weichselian glaciation lasted from 113 ka to ~ 13-9 ka in Iceland (Fronval

and Jansen, 1997). When the glaciation began to wane and the ice retreated, there was a

large, rapid pulse of magmatism resulting from isostatic rebound and decompression

melting. Therefore, a decrease in melt production in the WVZ related to the ridge

relocation described in the previous section may not be observable in the postglacial

Holocene because the signal would be masked by the sudden, extreme increase in melt

production associated with deglaciation (Sinton et al., 2005). Immediately following the

rapid removal of ice following the last glacial maximum, a spike in decompression

melting spawned abundant and voluminous volcanism, which tapered off abruptly after

7

only one to two thousand years (Jull and McKenzie, 1996; Maclennan et al., 2002). This

phenomenon has been documented on the Reykjanes Peninsula (Guðmundsson, 1986),

the WVZ (Sinton et al., 2005) and in all the other volcanic zones with production rates up

to 30 times that of equilibrium melting (Maclennan et al., 2002). This phenomenon has

also given rise to markedly different compositions in lavas erupted in interglacial periods

and those erupted immediately following deglaciation (Gee et al., 1998b; Maclennan et

al., 2002; Sinton et al., 2005).

1.3 Regional geochemical overview

As an ocean island on a mid-ocean ridge, Iceland is dominantly (although by no

means entirely) basaltic in composition. Geochemical differences have been recognized

globally between mid-ocean ridge basalts (MORB) and ocean island basalts (OIB; Sun

and McDonough, 1989), but these differences were not initially understood to have

originated from separate mantle sources. Schilling (1973) invoked multiple mantle

sources to explain chemical variation in Iceland. He noted a gradual depletion in light

rare earth elements (REEs) moving southwest along the Reykjanes Ridge away from

Icelandic plume. Large ion lithophile elements (LILEs) (such as K, Rb, Cs, Ba, U, and

Th) and light to intermediate REE ratios (notably La/Sm) in Reykjanes Ridge basalts

followed a similar gradational trend (Figure 1.2). Not only did Schilling (1973) suggest a

more primordial source beneath Iceland, but he used this evidence to argue for a two

component mixing model between melts from both MORB and a plume (OIB) source in

the mantle.

8

Contemporaneously, Jakobsson (1972) recognized a geographic correlation

between variations in major element compositions across Iceland. He proposed three

igneous rock series and linked them to their proximity to the MAR rift zones. These

remain the defining rock series of Iceland today and are correlated to tectonic

environment (Sigmarsson and Steinthórsson, 2007; Jakobsson et al., 2008). Alkali and

transitional alkali basalts are confined to the off-rift volcanic zones and near the south tip

of the propagating EVZ rift. Tholeiites, the most common of the basaltic rock series, are

found along the neovolcanic rift zones. These rock series correlate with the melt fraction

and volume, with tholeiites representing greater melt fraction and alkali basalts indicating

small batches of more localized melt. These major element compositions are probably

controlled more by tectonic environment and melting regime rather than variable mantle

sources.

A great deal of research has gone into determining the number and compositions

of different mantle domains contributing to Icelandic magmas. Among trace elements,

niobium seems to be the most robust tracer of primordial mantle contributions because

Nb is depleted in both depleted MORB-source upper mantle and in the crust relative to

primitive mantle (McKenzie and O’Nions, 1991; Fitton et al., 1997). Elevated Nb/Zr

ratios (≥0.08) in Iceland and at other North Atlantic hotspots relative to MAR ridge

segments (≤0.05) is consistent with theoretical Nb systematics (Figure 1.2; Fitton et al.,

1997; Hanan et al., 2000).

Helium isotope ratios are also broadly considered to be a robust tracer of a

primitive mantle component. The primordial mantle theoretically has a higher 3He/4He

ratio than degassed upper mantle because 4He will be replenished in a degassed mantle by

9

alpha decay of radionuclides, whereas 3He will not, driving 3He/4He to lower values over

time. Ratios are typically expressed relative to the atmospheric ratio (R/Ra) and it has

long been recognized that Iceland has R/Ra, as much as three times greater (R/Ra = 24)

than MORB and that these values are typically higher close to the center of the plume in

east-central Iceland, although this distribution is non-uniform (Condomines et al., 1983;

Kurz et al., 1985).

Indeed, the increasing consensus in the literature favors the mixing of multiple (2-

4) source domains within the plume based on Sr-Nd-Pb-Hf isotope systematics. Some

propose the range of compositions of Icelandic lavas arises without contribution from a

MORB source and that the plume contains at least one depleted (unradiogenic Pb,

radiogenic Nd, and moderate to unradiogenic Sr) component (Chauvel and Hémond,

2000; Thirlwall et al., 2004; Kokfelt et al., 2006). Rather than have a mid-ocean ridge

contaminated by a plume source, they posit that the plume itself is responsible for

heterogeneity in Iceland and instead, it contaminates the surrounding MAR. Regardless

of whether there are contributions from a MORB source (Hanan and Schilling, 1997;

Hanan et al., 2000) or not, most recent studies converge on the involvement of ancient

recycled oceanic crust or lithosphere (Chauvel and Hémond, 2000; Thirlwall et al., 2004;

Kokfelt et al., 2006; Peate et al., 2010; Koornneef et al., 2012).

Two dominant models for the origins of Icelandic magmas are best summarized

by Manning and Thirlwall (2014). The first results from the melting of a zoned plume or

one that contains veins or blobs of fertile material within depleted lithologies (Chauvel

and Hémond, 2000; Kokfelt et al., 2006; Koornneef et al., 2012). Either fertile pyroxenite

is incorporated into ambient peridotite or the structure of recycled oceanic lithosphere is

10

preserved, giving rise to bilaterally zoned plume with melts from both fertile basaltic and

refractory gabbroic compositions. The second model advocates convective mixing

between numerous, isotopically distinct sources within the plume (Thirlwall et al., 2004;

Peate et al., 2009; Peate et al., 2010). Excluding Öræfajökull, which requires the melting

of minor amounts of pelagic sediment (Manning and Thirlwall, 2014), the Iceland array

in 87Sr/86Sr and 143Nd/144Nd supports a binary mixing model. Lead isotope work has

provided the most robust evidence for a third enriched mantle source (Sigmarsson &

Steinthórsson 2007, and references therein) although Koornneef et al. (2012) argue that

one dimensional polybaric melting models with mixing of melts from two sources at

different depths can give rise to the full range of isotope compositions.

Oxygen isotopes in Icelandic lavas are commonly up to 2‰ lighter than mantle

δ18O values. This observation has also been used as evidence for a recycled oceanic

crustal component interspersed in the mantle below Iceland (Kokfelt et al., 2006).

Sigmarsson & Steinthórsson (2007), on the other hand, argue that contributions from the

altered Icelandic crust to magmatic systems may limit the use of oxygen isotopes to

identify mantle sources. Assimilation of altered basalt or hyaloclastite may drive

Icelandic lavas to δ18O several ‰ lighter than mantle values based on observed extreme

negative δ18O (<-10‰) in hydrothermal systems (Muehlenbachs et al., 1974; Hattori and

Muehlenbachs, 1982) and more moderate depletions in δ18O in the ubiquitous

hyaloclastite formations in the EVZ (Bindeman et al., 2008).

11

1.4. The Reykjanes Peninsula and the Brennisteinsfjöll volcanic system

1.4.1 Reykjanes Peninsula tectonics and volcanism

The Reykjanes Peninsula, in southwest Iceland is a oblique transtensional plate

boundary that arcs onshore to Iceland from the Reykjanes Ridge. It connects the offshore

Reykjanes Ridge with the onshore Western Volcanic Zone (a ridge) and the South

Iceland Seismic Zone (a “transform”) at a triple junction at Hengill. It consists of 4 NE-

SW trending en echelon volcanic systems that form the plate boundary (Figure 1.3).

Pioneering microseismic observations showed that both strike-slip and normal faults are

present across the peninsula (Klein et al., 1973; Klein et al., 1977), which was supported

by geodetic data that indicated both left-lateral and extensional motion (Brander et al.,

1976). These studies demonstrated that the RP was not simply a transform plate boundary

connecting the Reykjanes Ridge to the WVZ, but rather an extensional transform zone or

what is sometimes referred to as a leaky transform (Taylor et al., 1994).

A variety of techniques have been implemented to understand the distribution of

faults, volcanic fissures, and fractures on the Reykjanes that partition strain. These

methods include the use of seismic data (Keiding et al., 2009), GPS geodesy (Hreinsdóttir

et al., 2001; Clifton et al., 2002; Árnadóttir et al., 2006; Keiding et al., 2008), InSAR

(Clifton et al., 2002), analog models (Clifton and Schlische, 2003), structural analysis

(Villemin and Bergerat, 2013), and high resolution DEM mapping of fissures (Clifton

and Kattenhorn, 2006).

Geodetic GPS work focuses on determining how NUVEL-1A spreading rates of

19-20 mm/yr between the North American and Eurasian plate are partitioned on the

12

oblique divergent Reykjanes Peninsula. From 1993-1998 GPS velocities, Hreinsdóttir et

al. (2001) observed 16.8 ± 0.9 mm/yr of left-lateral motion on the RP, but not the

expected extensional component. They postulate that extension is accommodated during

episodic rifting events that are thought to be cyclical with a ~1000 year periodicity

(Clifton and Kattenhorn, 2006; Keiding et al., 2008). The last eruption on the RP

occurred in the 13th century and the peninsula is presently in a quiescent period.

However, recent work by Árnadóttir et al. (2006) models 1992-2000 GPS velocities with

a best-fit locking depth of 8 km, and 19 mm/yr of left-lateral transcurrent motion, and ~9

mm/yr of extension at depth.

The motion and orientation of the plate boundary at depth gives rise to more

complex deformation in the brittle regime and expression of structures at the surface.

Clifton and Katterhorn (2006) use high resolution digital elevation maps (DEM) to map

the surface expression of faults, fissures, and fractures on the Reykjanes. They find that

the major strike-slip faults are oriented N-S, rotated 80-90° clockwise from the

orientation of plate motion, and are primarily to the SE of the volcanic systems. The

normal faults, on the other hand, are oriented NE-SW, parallel or sub-parallel to the trend

of eruptive fissures and rotated ~60-70° counterclockwise to the direction of plate

motion. North trending right-lateral strike-slip faults are en echelon and accommodate the

E-W sinistral shear across the plate boundary; this style of faulting is referred to as

bookshelf faulting and is most developed on the eastern Reykjanes. This style of faulting

is characteristic of the SISZ to the east of the triple junction at Hengill (Bergerat and

Angelier, 2003; Clifton and Einarsson, 2005; Clifton and Kattenhorn, 2006), however

there is no observed extension in the SISZ. The commonality between the well-developed

13

N-S oriented dextral faulting in the SISZ and the eastern RP suggests that the volcanic

systems of Brennisteinsfjöll and Hengill are governed by a tectonic regime intermediate

between that of the RP and the amagmatic SISZ.

1.4.2 Reykjanes Peninsula lava chemistry

Except for rare silicic rocks from the Hengill volcanic system, the Reykjanes

Peninsula is uniformly basaltic in composition. High MgO lavas on the RP fall broadly

into two geochemical groups: depleted and enriched basalts (Gee et al., 1998a; Thirlwall

et al., 2004; Peate et al., 2009). Depleted lavas have Nb/Zr < 0.08, 87Sr/86Sr < 0.70311,

and 143Nd/144Nd > 0.51307 (Gee et al., 1998a). Peate et al. (2009) sampled all of the

historic lavas on the Reykjanes (since the settlement of the island circa 871 C.E.). These

are uniformly enriched, but are subdivided into “less enriched” (Nb/Zr 0.10-0.13, La/YbN

1.3-2.0) and “more enriched” (Figure 1.4; Nb/Zr 0.13-0.14, La/YbN 2.2-2.6, 206Pb/204Pb

>18.8).

While many authors have focused on depleted picrites because they are the most

primitive compositions on the RP and therefore perhaps most accurately reflect primary

mantle melts, these are not representative and occur in rare, small volume flows

(Jakobsson et al., 1978; Peate et al., 2009). In fact, these lavas are constrained primarily

to 9-13 ka, during and immediately following deglaciation. The local stress state on

magma bodies is likely to have fluctuated during glacial unloading, which may have

caused eruptions to occur with increased frequency. Gee et al (1998b) argue that reduced

residence times immediately following deglaciation prevent RP lavas from acquiring

14

enriched trace element and isotope compositions from crustal contamination.

Additionally, they posit that shorter residence times in the crust of smaller volumes of

magma diminished the potential for magma mixing to occur, thereby preserving the

depleted melt.

While crustal assimilation has been demonstrated to be an important magmatic

process on the Reykjanes Peninsula (Gee et al., 1998a), the composition of these lavas

has been reinterpreted to reflect binary mixing of melts derived from depleted and

enriched components in the Icelandic plume. The enriched component dominates,

comprising 50-90% of the mix for majority of RP lavas (Thirlwall et al., 2004; Peate et

al., 2009) and the preservation of the depleted melts may be facilitated by the reduced

residence times and decreased probability of magma mixing (Gee et al., 1998b).

Nevertheless, there is a strong correlation between lava chemistry and age over the last

20 ka, with modern compositions at or approaching those before the deglaciation in

southwest Iceland (Figure 1.5).

Rifting seems to be confined to one part of the northern Reykjanes Ridge or the

RP at any given time and the estimated recurrence interval of these rifting episodes and

associated volcanism is ~1000 years (Clifton and Kattenhorn, 2006; Keiding et al., 2008).

The historic lavas from 940-1340 C.E. capture one of these intervals, within which

enrichment in lavas increases to the east (Peate et al., 2009). These authors explicitly

refrain from suggesting a causation for this correlation by stating that the full range of

enriched values has been expressed on the Reykjanes over the last 9 ka (Gee et al., 2000).

However, no other authors explore coincident rifting episodes across volcanic systems on

the Reykjanes, so this correlation cannot be yet be dismissed as chance. Crustal thickness,

15

subtle variations in tectonics and/or magmatic processes, and proximity to the plume

center in the EVZ among other factors cannot yet be ruled out.

16

Figure 1.1. Map of Iceland and its neovolcanic zones. These zone trace the plate boundary between North America (NA) and Eurasia (EU). KR =

Kolbeinsey Ridge; TFZ = Tjörnes Fracture Zone; NVZ = Northern Volcanic Zone; EVZ = Eastern Volcanic Zone; MIB = Mid-Iceland Belt; WVZ

= Western Volcanic Zone; RP = Reykjanes Peninsula; RR = Reykjanes Ridge; R = Reykjanes; K = Krísuvík; BSF = Brennisteinsfjöll; H = Hekla;

VME = Vestmannaeyjar. Volcanic zones are composed of many individual volcanic systems (Einarsson and Sæmundsson, 1987). Contours mark

the approximate depth to the Moho as determined by Darbyshire et al. (2000).

17

Figure 1.2. Variations along the Mid-Atlantic Ridge in La/Sm and Nb/Zr. Values are <1 for La/Sm on the

South Reykjanes Ridge (SRR) and the Kolbeinsey Ridge (KR) and <0.05 for Nb/Zr. These ratios are

greater than 1.5 and 0.08 for mainland Iceland (I). The North Reykjanes Ridge (NRR) is transitional

between ambient MORB compositions and those in lavas derived from the Icelandic plume. Data from

(Hanan et al., 2000).

18

Figure 1.3. Map of the Reykjanes Peninsula showing the four fissure systems (shaded grey). From east to west they are the Reykjanes, Krísuvík,

Brennisteinsfjöll, and Hengill volcanic systems. Thríhnúkagígur is located in Brennisteinsfjöll. Reykjavík and Thríhnúkagígur are shown in black

stars. Outlines of the volcanic systems are from Einarsson & Sæmundsson (1987).

19

Figure 1.4. Enriched historic Reykjanes Peninsula lavas plotted on Nb/Zr vs. 87Sr/86Sr (a) and Nb/Zr vs.

143Nd/144Nd. Data from Peate et al. (2009) fall into “more enriched” and “less enriched” categories.

More enriched sample correlate with lavas from Brennisteinsfjöll, while less enriched data are from the

eastern Reykjanes Peninsula fissure swarms of Reykjanes and Krísuvík (b).

20

Figure 1.5. Changes in lava chemistry on the Reykjanes Peninsula through time from Gee et al. (1998b).

Plots of lava age vs. Mg# (a), 87Sr/86Sr (b), Nb/Zr (c), and 143Nd/144Nd (d). The period during and

immediately following deglaciation (13-9 ka) produced more primitive, depleted magmas with higher

Mg# and 87Sr/86Sr and lower Nb/Zr ratios and 143Nd/144Nd.

21

References

Árnadóttir T., Jiang W., Feigl K. L., Geirsson H. and Sturkell E. (2006) Kinematic models of

plate boundary deformation in southwest Iceland derived from GPS observations. J.

Geophys. Res. 111, 1–16.

Bergerat F. and Angelier J. (2003) Mechanical behaviour of the Árnes and Hestfjall Faults of the

June 2000 earthquakes in Southern Iceland: Inferences from surface traces and tectonic

model. J. Struct. Geol. 25, 1507–1523.

Bindeman I., Gurenko A., Sigmarsson O. and Chaussidon M. (2008) Oxygen isotope

heterogeneity and disequilibria of olivine crystals in large volume Holocene basalts from

Iceland: Evidence for magmatic digestion and erosion of Pleistocene hyaloclastites.

Geochim. Cosmochim. Acta 72, 4397–4420.

Bjarnason I. T. (2008) An Iceland hotspot saga. Jökull, 3–16.

Bjarnason I. T. and Schmeling H. (2009) The lithosphere and asthenosphere of the Iceland

hotspot from surface waves. Geophys. J. Int. 178, 394–418.

Bjarnason I. T., Wolfe C. J. and Solomon S. C. (1996) Initial results from the ICEMELT

experiment : Body-wave delay times and shear-wave splitting across Iceland. Geophys. Res.

Lett. 23, 459–462.

Björnsson A., Eysteinsson H. and Beblo M. (2005) Crustal formation and magma genesis beneath

Iceland: Magnetotelluric constraints. Geol. Soc. Am. Spec. Pap. 388, 665–686.

Brander J. L., Mason R. G. and Calvert R. W. (1976) Precise distance measurements in Iceland.

Tectonophysics 31, 193–206.

Brandsdóttir B. and Menke W. H. (2008) The seismic structure of Iceland. Jökull 58, 17–34.

Brounce M., Feineman M., LaFemina P. C. and Gurenko A. (2012) Insights into crustal

assimilation by Icelandic basalts from boron isotopes in melt inclusions from the 1783-1784

Lakagígar eruption. Geochim. Cosmochim. Acta 94, 164–180.

Chauvel C. and Hémond C. (2000) Melting of a complete section of recycled oceanic crust: Trace

element and Pb isotopic evidence from Iceland. Geochemistry, Geophys. Geosystems 1, n/a–

n/a.

Clifton A. E. and Einarsson P. (2005) Styles of surface rupture accompanying the June 17 and 21,

2000 earthquakes in the South Iceland Seismic Zone. Tectonophysics 396, 141–159.

Clifton A. E. and Kattenhorn S. A. (2006) Structural architecture of a highly oblique divergent

plate boundary segment. Tectonophysics 419, 27–40.

22

Clifton A. E. and Schlische R. W. (2003) Fracture populations on the Reykjanes Peninsula,

Iceland: Comparison with experimental clay models of oblique rifting. J. Geophys. Res.

108.

Clifton A. E., Sigmundsson F., Feigl K. L., Gunnar Guðmundsson and Árnadóttir T. (2002)

Surface effects of faulting and deformation resulting from magma accumulation at the

Hengill triple junction, SW Iceland, 1994-1998. J. Volcanol. Geotherm. Res. 115, 233–255.

Condomines M., Muehlenbachs K., O’Nions R. K., Óskarsson N. and Oxburgh E. R. (1983)

Helium, oxygen, strontium and neodynium isotopic relationships in Icelandic volcanics.

Earth Planet. Sci. Lett. 66, 125–136.

Darbyshire F. A., Bjarnason I. T. and White R. S. (1998) Crustal structure above the Iceland

mantle plume imaged by the ICEMELT refraction profile. Geophys. J. Int. 135, 1131–1149.

Darbyshire F. A., White R. S. and Priestley K. F. (2000) Structure of the crust and uppermost

mantle of Iceland from a combined seismic and gravity study. Earth Planet. Sci. Lett. 181,

409–428.

DeMets C., Gordon R. G. and Argus D. F. (2010) Geologically current plate motions. Geophys. J.

Int. 181, 1–80.

Einarsson P. and Sæmundsson K. (1987) Earthquake epicenters 1982-1985 and volcnaic systems

in Iceland (map). In I hlutarins eðli: Festschrift for Thorbjorn Sigurgeirsson (ed. Þ.

Sigfússon).

Fitton J. G., Saunders A. D., Norry M. J., Hardarson B. S. and Taylor R. N. (1997) Thermal and

chemical structure of the Iceland plume. Earth Planet. Sci. Lett. 153, 197–208.

Foulger G. R. and Anderson D. L. (2005) A cool model for the Iceland hotspot. J. Volcanol.

Geotherm. Res. 141, 1–22.

Foulger G. R., Pritchard M. J., Julian B. R., Evans J. R., Allen R. M., Nolet G., Morgan W. J.,

Bergsson B. H., Erlendsson P., Jakobsdóttir S., Ragnarsson S., Stefansson R. and Vogfjo K.

(2000) The seismic anomaly beneath Iceland extends down to the mantle transition zone and

no deeper. Geophys. J. Int. 142, F1–F5.

Fronval T. and Jansen E. (1997) Eemian and Early Weichselian (140–60 ka) Paleoceanography

and paleoclimate in the Nordic Seas with comparisons to Holocene conditions.

Paleoceanography 12, 443.

Gee M. A. M., Taylor R. N., Thirlwall M. F. and Murton B. J. (2000) Axial magma reservoirs

located by variation in lava chemistry along Iceland’s mid-ocean ridge. Geology 28, 699–

702.

Gee M. A. M., Taylor R. N., Thirlwall M. F. and Murton B. J. (1998) Glacioisostacy controls

chemical and isotopic characteristics of tholeiites from the Reykjanes Peninsula, SW

Iceland. Earth Planet. Sci. Lett. 164, 1–5.

23

Gee M. A. M., Thirlwall M. F., Taylor R. N., Lowry D. and Murton B. J. (1998) Crustal

Processes: Major Controls on Reykjanes Peninsula Lava Chemistry, SW Iceland. J. Petrol.

39, 819–839.

Geirsson H., Bennett R. A., LaFemina P. C., Árnadóttir T., Sturkell E., Sigmundsson F., Travis

M., Schmidt P., Lund B. and Hreinsdóttir S. (2012) Volcano deformation at active plate

boundaries: Deep magma accumulation at Hekla volcano and plate boundary deformation in

south Iceland. J. Geophys. Res. 117, B11409.

Guðmundsson A. (1986) Mechanical Aspects of Postglacial Volcanism and Tectonics of the

Reykjanes Peninsula, Southwest Iceland. J. Geophys. Res. 91, 12711–12721.

Hanan B. B., Blichert-Toft J., Kingsley R. and Schilling J.-G. (2000) Depleted Iceland mantle

plume geochemical signature: Artifact of multicomponent mixing? Geochemistry, Geophys.

Geosystems 1.

Hanan B. B. and Schilling J.-G. (1997) The dynamic evolution of the Iceland mantle plume: the

lead isotope perspective. Earth Planet. Sci. Lett. 151, 43–60.

Hardarson B. S., Fitton J. G., Ellam R. M. and Pringle M. S. (1997) Rift relocation - a

geochemical and geochronological investigation of a palaeo-rift in northwest Iceland. Earth

Planet. Sci. Lett. 153, 181–196.

Hattori K. and Muehlenbachs K. (1982) Oxygen Isotope Ratios of the Icelandic Crust. J.

Geophys. Res. 87, 6559–6565.

Hreinsdóttir S., Einarsson P. and Sigmundsson F. (2001) Crustal deformation at the oblique

spreading Reykjanes Peninsula, SW Iceland: GPS measurements from 1993 to 1998. J.

Geodyn. 106, 13803–13816.

Jakobsson S. P. (1972) Chemistry and distribution pattern of Recent basaltic rocks in Iceland.

Lithos 5, 365–386.

Jakobsson S. P., Jónasson K. and Sigurdsson I. A. (2008) The three igneous rock series of

Iceland. Jökull 58, 117–138.

Jakobsson S. P., Jonsson J. and Shido F. (1978) Petrology of the Western Reykjanes, Peninsula.

J. Petrol. 19, 669–705.

Jull M. and Mckenzie D. (1996) The effect of deglaciation on mantle melting beneath Iceland. J.

Geophys. Res. 101, 21815–21828.

Keiding M., Árnadóttir T., Sturkell E., Geirsson H. and Lund B. (2008) Strain accumulation

along an oblique plate boundary: The Reykjanes Peninsula, southwest Iceland. Geophys. J.

Int. 172, 861–872.

Keiding M., Lund B. and Árnadóttir T. (2009) Earthquakes, stress, and strain along an obliquely

divergent plate boundary: Reykjanes Peninsula, southwest Iceland. J. Geophys. Res. Solid

Earth 114, 1–16.

24

Klein F. W., Einarsson P. and Wyss M. (1973) Microearthquakes on the Mid-Atlantic Plate

Boundary on the Reykjanes Peninsula in Iceland. J. Geophys. Res. 78, 5084.

Klein F. W., Einarsson P. and Wyss M. (1977) The Reykjanes Peninsula, Iceland, Earthquake

Swarm of September 1972 and Its Tectonic Significance. J. Geophys. Res. 82, 865–888.

Kokfelt T. F., Hoernle K., Hauff F., Fiebig J., Werner R. and Garbe-Schönberg D. (2006)

Combined Trace Element and Pb-Nd-Sr-O Isotope Evidence for Recycled Oceanic Crust

(Upper and Lower) in the Iceland Mantle Plume. J. Petrol. 47, 1705–1749.

Koornneef J. M., Stracke A., Bourdon B., Meier M. A., Jochum K. P., Stoll B. and Grönvold K.

(2012) Melting of a Two-component Source beneath Iceland. J. Petrol. 53, 127–157.

Kurz M. D., Meyer P. S. and Sigurdsson H. (1985) Helium isotopic systematics within the

neovolcanic zones of Iceland. Earth Planet. Sci. Lett. 74, 291–305.

LaFemina P. C., Dixon T. H., Malservisi R., Árnadóttir T., Sturkell E., Sigmundsson F. and

Einarsson P. (2005) Geodetic GPS measurements in south Iceland: Strain accumulation and

partitioning in a propagating ridge system. J. Geophys. Res. 110, B11405.

Larsen L. M., Pedersen A. K., Pedersen G. K. and Piasecki S. (1992) Timing and duration of

early Tertiary volcanism in the North Atlantic: new evidence from West Greenland.

Magmat. causes Cont. Break. 68, 321–333.

Maclennan J., Jull M., McKenzie D., Slater L. and Grönvold K. (2002) The link between

volcanism and deglaciation in Iceland. Geochemistry, Geophys. Geosystems 3, 1–25.

Manning C. J. and Thirlwall M. F. (2014) Isotopic evidence for interaction between Öræfajökull

mantle and the Eastern Rift Zone, Iceland. Contrib. to Mineral. Petrol. 167, 959.

McKenzie D. and O’Nions R. K. (1991) Partial Melt Distributions from Inversion of Rare Earth

Element Concentrations. J. Petrol. 32, 1021–1091.

Mittelstaedt E., Ito G. and Behn M. D. (2008) Mid-ocean ridge jumps associated with hotspot

magmatism. Earth Planet. Sci. Lett. 266, 256–270.

Morgan W. J. (1971) Convection Plumes in the Lower Mantle. Nature 230, 42–43.

Muehlenbachs K., Anderson A. T. and Sigvaldason G. E. (1974) Low-O18 basalts from Iceland.

Geochim. Cosmochim. Acta 38, 577–588.

Nunns A. G., Talwani M., Lorentzen G. R., Vogt P. R., Sigurgeirsson T., Kristjánsson L., Larsen

H. C. and Voppel D. (1983) Magnetic anamolies over Iceland and surrounding seas. In

Structure and Development of the Greenland-Scotland Ridge pp. 661–678.

Peate D. W., Baker J. A., Jakobsson S. P., Waight T. E., Kent A. J. R., Grassineau N. V. and

Skovgaard A. C. (2009) Historic magmatism on the Reykjanes Peninsula, Iceland: a snap-

shot of melt generation at a ridge segment. Contrib. to Mineral. Petrol. 157, 359–382.

25

Peate D. W., Breddam K., Baker J. A., Kurz M. D., Barker A. K., Prestvik T., Grassineau N. V.

and Skovgaard A. C. (2010) Compositional Characteristics and Spatial Distribution of

Enriched Icelandic Mantle Components. J. Petrol. 51, 1447–1475.

Saunders A. D., Fitton J. G., Kerr A. C., Norry M. J. and Kent R. W. (1997) The North Atlantic

Igneous Province. In Large Igneous Provinces: Continental, Oceanic, and Planetary Flood

Volcanism pp. 45–93.

Schilling J.-G. (1973) Iceland Mantle Plume: Geochemical Study of Reykjanes Ridge. Nature

242, 565–571.

Sigmarsson O. and Steinthórsson S. (2007) Origin of Icelandic basalts: A review of their

petrology and geochemistry. J. Geodyn. 43, 87–100.

Sinton J., Grönvold K. and Sæmundsson K. (2005) Postglacial eruptive history of the Western

Volcanic Zone, Iceland. Geochemistry, Geophys. Geosystems 6, n/a–n/a.

Sun S. -s. and McDonough W. F. (1989) Chemical and isotopic systematics of oceanic basalts:

implications for mantle composition and processes. Geol. Soc. London, Spec. Publ. 42, 313–

345.

Taylor B., Crook K. and Sinton J. (1994) Extensional transform zones and oblique spreading

centers. J. Geophys. Res. 99, 19707–19718.

Thirlwall M. F., Gee M. A. M., Taylor R. N. and Murton B. J. (2004) Mantle components in

Iceland and adjacent ridges investigated using double-spike Pb isotope ratios. Geochim.

Cosmochim. Acta 68, 361–386.

Villemin T. and Bergerat F. (2013) From surface fault traces to a fault growth model: The Vogar

Fissure Swarm of the Reykjanes Peninsula, Southwest Iceland. J. Struct. Geol. 51, 38–51.

Walters R. L., Jones S. M. and Maclennan J. (2013) Renewed melting at the abandoned Húnafloí

Rift, northern Iceland, caused by plume pulsing. Earth Planet. Sci. Lett. 377-378, 227–238.

Weir N. R. W., White R. S., Brandsdóttir B., Einarsson P., Shimamura H. and Shiobara H. (2001)

Crustal structure of the northern Reykjanes Ridge and Reykjanes Peninsula, southwest

Iceland. J. Geophys. Res. 106, 6347.

Wilson J. T. (1963) Continental Drift. Sci. Am. 208, 86–103.

Wolfe C. J., Bjarnason I. T., VanDecar J. C. and Solomon S. C. (1997) Seismic structure of the

Iceland mantle plume. Nature 385, 245–247.

26

Chapter 2

Inside the Volcano: Rapid entrainment of unconsolidated crust during a

monogenetic fissure eruption

2.1 Introduction

Small volume basaltic eruptions (<1 km3) produce tephra and scoria cones and/or

lava fields that are among the most abundant subaerial volcanic landforms on Earth

(Wood, 1980; Keating et al., 2007). As commonplace geologic features, it is important to

understand how their structure and development are controlled by the preexisting

structures relative to the far field stresses that typically control dike propagation and

orientation. These eruptions occur in a variety settings including monogenetic fields, at

the sites of fissure eruptions, and on the flanks of stratovolcanoes and shield volcanoes.

Recent rifting events have been well-observed and well-imaged geodetically (Wright et

al., 2006; Calais et al., 2008; Biggs et al., 2013; Sigmundsson et al., 2015), but are not

able to fully resolve the geometry and development of the uppermost eruptive conduit.

Data from older, eroded fissure systems have been described in detail (Németh and

White, 2003; Keating et al., 2007; Genareau et al., 2010; Valentine, 2012; Harp and

Valentine, 2015; Németh and Kereszturi, 2015). These suggest that decrease in the

vertical stress field in the very uppermost crust allows dikes to widen in the upper 50-150

m (Keating et al., 2007; Valentine, 2012). Yet, most observations of the internal structure

of monogenetic cones are limited to eroded systems at the surface of the Earth. Rarely is

the full length of the shallow magma plumbing observed in entirety.

Thríhnúkagígur is a monogenetic, small volume basaltic fissure eruption on the

Reykjanes Peninsula. It is unique in Iceland, and perhaps in the world, in that a large cave

27

directly below the main cone exposes a cross-section of the uppermost magma plumbing

system, as well as the regional stratigraphy. This vertically oriented lung-shaped cave

offers a unique a view into some non-unique magmatic and eruptive processes. First, the

uppermost ~150 m of architecture provides insights into the eruptive process and the

behavior of dikes in the shallow subsurface. Second, the remnants of a buried tephra cone

near the bottom of the cave provides insights to one way a crustal contaminant may be

added to mantle-derived magmas. The buried tephra cone is evaluated for both its role in

focusing the flux of magma to a single vent at the surface and as a crustal material that

could have been entrained by the erupting magma, thereby creating a vacancy below the

main cone.

2.2 Field descriptions

2.2.1 Surface geology

Thríhnúkar is a NNE-SSW trending fissure system in the Brennisteinsfjöll

volcanic system. Its name is derived from the three peaks, Vesturhnúkur, Miðhnúkur, and

Thríhnúkagígur that mark different eruptive episodes (Figure 2.1; Sæmundsson, 2006).

Vesturhnúkur, the oldest and westernmost of the peaks, sits on the edge of a plateau that

descends steeply to the northwest. It is associated temporally with the late Weichselian

glaciation (Sæmundsson et al., 2010) and consists primarily of basaltic hyaloclastite. Thin

dikes are pervasive on the western slope of the hyaloclastite and three dolerite plugs

occur at the summit, representing potential eruptive centers.

28

Miðhnúkur is a basaltic spatter cone between Vesturhnúkur and Thríhnúkagígur

that is estimated to be approximately 5 ka in age (Sæmundsson, 2006; and references

therein). Flows from Miðhnúkur (H-145; Figure 2.1) are the most extensive of the three

Thríhnúkar eruptions with some outcrops appearing up to 6.5 km away from the main

cone (Sæmundsson et al., 2010). Most of the flow extends off the plateau to the west and

northwest. Its full extent is unknown because it has been largely buried by subsequent

lava flows from other eruptive fissures. A 300 meter long depression formed east of the

Miðhnúkur cone during the event, and looks like a collapsed lava tube. It is elongate

parallel to the strike of the fissure and at least partially filled with lavas from the

Miðhnúkar eruption.

The eruption of Thríhnúkagígur is estimated to be 3.5 ka in age (Sæmundsson,

2006). Soil profiles overlying Thríhnúkagígur lavas included Katla tephras from >2500

bp (Sæmundsson, 2006; Sæmundsson, 2008).The easternmost of the Thríhnúkur peaks,

its main cone is an amalgamation of two vents from a fissure eruption. Smaller disparate

spatter cones associated with the fissure eruption occur up to 1.5 km from the main cone

(Figure 2.1). Thríhnúkagígur lavas can be distinguished in the field by the presence of

macroscopic plagioclase phenocrysts as opposed to olivine in the Miðhnúkar lavas. The

main flow from Thríhnúkagígur (H-146) covers roughly 0.34 km2 (Figure 2.1). Three

additional small lava fields are present along the fissure and were fed by small spatter

cones. One is immediately adjacent to the main cone to the SSW and the other two lava

fields are 1.5 km to the NNE on the flank of the plateau (Figure 2.1). These three flows

total to just under 0.1 km2 in area. The one proximal to the main cone is only ~0.015 km2

in area.

29

Lavas flows are relatively thin, about 15-30 cm in height. The flow tops are

vesicle-poor, but become increasing vesicular towards the base of the flow where

individual vesicles are up to several centimeters in diameter and become more

interconnected. Lavas from the main cone flowed primarily to the north and east, but also

filled the depression to its west that had formed originally during the Miðhnúkar eruption.

The depression filled to a height of about 1-1.5 m leaving behind a terrace (bathtub ring)

after it drained to the east by a lava tube.

The main spatter cone is relatively symmetrical, but elongate along the azimuth of

the fissure. The summit consists of a pair of similarly sized vents at the north and south

ends of the cone. The southern vent remained open, whereas the northern vent became

plugged towards the end of the eruption. The cone formed from the coalescence of the

fissure into a couple of main vents during a Strombolian eruption with low viscosity, gas-

rich lavas (Figure 2.2). Pele’s tears and oxidized scoriaceous vent spatter of various

colors around and at the top of the cone support this eruptive style. There are also several

tens of vesicle-poor 2-5 cm rheomorphic flows draining back into the vent (Figure 2.3).

These thin flows back into the vent contrast the surrounding highly vesicular H-146

flows. The vent itself is ~8 m in diameter at the surface, but narrows rapidly with

increasing depth, in places less than 3 by 2 meters, elongated along the fissure azimuth

reflecting the overall cone structure (Figure 2.2).

2.2.2 Subsurface Geology

The eruptive conduit provides the only entrance to the subsurface of

Thríhnúkagígur (Figure 2.4). The narrow cylindrical section of the conduit extends

30

vertically 60 m down from the summit. This section does not expose much of the

subsurface stratigraphy as it is coated by flowback of lava into the conduit. Flowback is

characterized by rivulet textures (Figure 2.5) and stalactites of lava (Figure 2.6) that are a

few centimeters to tens of cm in scale. Part way down the conduit there is a narrow,

continuous tunnel that diverts off and reconnects to the main conduit in the shape of a

teacup handle.

Further down the conduit from the surface, the cave broadens and flowback

features no longer coat the cave walls. As the cylindrical conduit opens up into the main,

lung-shaped hollow of the cave, it transitions from a focused oblong tube into a 2 m wide

planar dike exposed in the cave walls. The majority of the exposed rock consists of sub-

horizontal basalt flows that have been altered to a mustard yellow (Figure 2.7). These

flows are commonly a few tens of cm in thickness, and are generally not more than a

meter thick.

Underlying the succession of lava flows is a buried pile of tephra, which has

previously been interpreted as a laterally continuous, gently dipping layer of hyaloclastite

that breaks the surface at Vesturhnúkar (Sæmundsson, 2006). This unconsolidated

material observed in the cave beneath Thríhnúkagígur appears to be a separate geologic

unit from the hyaloclastite. Instead, it is interpreted herein as the remnant of a subaerially

erupted tephra cone rather than a hyaloclastite ridge. The photo in Figure 2.8 shows the

contrast in physical characteristics between these two units. Although these two units

may represent different stages the hyaloclastite succession, rare bomb sags in the buried

tephra suggest a subaerial eruption. The buried is only exposed in the NW wall of the

cave and not on the SE wall, until deeper down in the cave. The contact with the

31

overlying lavas dips toward the NE and SW in wall, which is consistent with a tephra

cone that would dip in all directions from its highest point. The buried tephra cone is seen

interfingering with the overlying lava pile in the NW wall of the cave, but the slope of the

tephra remains relatively constant until it grades into a >1.5 m thick tephra sheet exposed

at the northern end of the cave (Figures 2.9, 2.10). Bombs and bomb sags in the tephra

apron suggest that the eruption that formed the tephra cone was subaerial. The constant

slope of the buried tephra cone and lack of reworking, other than minor interfingering

with subsequent lava flow, also suggests rapid burial by lava since the tephra is

unconsolidated. Because this relatively unconsolidated material is preserved at all, its

eruption and burial under ~120 m of lava flows likely occurred within an interglacial

period. Otherwise glaciers would have likely modified or removed the tephra cone.

The buried tephra primarily consists of unconsolidated, 1-5 mm glass shards with

rare macroscopic plagioclase crystals. Locally within the tephra pile pieces of scoria are

common, but no larger than 10 centimeters and rarely greater than 5 cm. The buried

tephra displays no major internal structures, but it is locally heterogeneous in size, color

and composition. For example, there is a 20 cm layer within the tephra that is well

consolidated with a quasi-foliated clay matrix around larger scoria clasts up to ~5 cm,

which suggests this tephra was reworked in the presence of water. Separately, in the

massive portions of the unit that lack evidence of reworking, there are zones that exhibit

1-2 cm wide spots that have been partially palagonitized.

The age of the buried tephra cone is unknown, but may be constrained by the

periodicity of magmatism and rifting on the Reykjanes Peninsula, which is thought to be

on the order of 1000 years (Sæmundsson et al., 2010). Lava flows in the cave are rarely

32

greater than 1 m thick. By assuming a rate of 1 meter of burial per 1000 years, and

accounting for the thickness of the lavas overlying the tephra apron (~120 m), an

approximate age of 100 ka is calculated. This assumes eruptions occurred in this location

during each rifting cycle and produced a lava flow thicker than the average flows

observed in the cave, so this is a minimum age estimate. The Eemian interglacial ended

approximately 113 ka with the onset of the Weichselian glaciation in northern Europe

(Fronval and Jansen, 1997), further suggesting that 100 ka is a minimum age.

Within the lava flows that lie stratigraphically above the top of the tephra cone in

the western wall of the cave there is a 5 m wide black to purple oval-shaped plug of

crystalline basalt with a large gas cavity in the center (Figure 2.11). It has an oxidized red

contact aureole and is interpreted to be an intrusion associated with the ~3.5 Ka

Thríhnúkagígur eruption. Although it has no surface expression, it is close enough to the

surface that it is probable that it vented to the surface and then was subsequently buried

during the formation of the main cone of Thríhnúkagígur. The plug itself is too high in

the ceiling to collect samples directly. Large boulders of oxidized, purple basalt beneath

the plug are interpreted as having fallen from the plugged conduit rather than the dike for

two reasons. First, some boulders have greater dimensions that the width of the dike.

Second, these boulders are not below the dike and are higher up on debris pile (the cave

floor) than any locations directly below the dike.

From the floor of the cave, the larger structure of the dike that fed the 3500 bp

eruption can be observed. The dike arcs across the ceiling tracing a vertical plane through

the center of the cave. It trends NNE-SSW, which is consistent with the regional

extensional fissures and crater rows of the Brennisteinsfjöll fissure swarm. The dike is

33

similar to the plugged conduit in that it is oxidized from black to purple and the baked

margins of country rock adjacent to the dike are red to a reddish-purple. About half way

down from the narrowing of the conduit to the floor of the cave, the dike has an en

echelon structure (Figure 2.7). Towards the floor of the cave on the northeast side, the

dike decreases in width to 0.5-1 m and develops a more complex structure. Further down

towards the floor on the northeast end of the cave, decimeter wide splays are developed.

In other segments, it separates into two thinner dikes then recombines into a single larger

dike. At the lowest exposure, the dike can be seen cross-cutting the tephra apron (Figure

2.9).

At the southwest end of the cave, a narrow gap between the debris pile and the

ceiling of the cave extends down along the strike of the dike. In one segment of the dike

there is a nearly 0.5 m long, 5 cm wide gas pocket that pinches out at both its ends

(Figure 2.11). Roughly 20 m down into this gap becomes a series of crawl spaces

between boulders. Here the dike can be seen crosscutting the tephra pile, but the structure

of the dike is difficult to trace because of the dike splays into many smaller dikes at this

depth and the exposure is much more limited. The best exposures of the dike are where it

fed vertical cylindrical pipes, or proto-conduits, that extend up into the buried tephra

(Figure 2.12). These proto-conduits have flowback features along their walls, suggesting

that their formation was a result of magmatic scouring of the poorly consolidated tephra.

These cylindrical cavities are typically immediately adjacent to the dike and range in

diameter from 0.25 to 2 m. The largest is mapped (Stefánsson, 1992; Sæmundsson, 2006)

and penetrates more than 30 m up into the overlying tephra (Figure 2.12).

34

2.3 Methods

Measuring the volume of the cave can help constrain the amount of tephra that

once filled the cave if it formed by the removal of unconsolidated buried tephra by an

erupting magma. Measuring the volume of the Thríhnúkagígur spatter cone and

estimating the lava volumes and their density can constrain the dense rock equivalent

(DRE) of the erupted products, which can help to constrain the relative proportion of

tephra that could have been entrained and incorporated into the lavas.

2.3.1 Light Detection and Ranging (LiDAR)

The cave at Thríhnúkagígur was scanned to gain an understanding of its overall

shape and structure and to relate that to the dynamics of the fissure eruption to which it is

likely related. The scans also serve the purpose of giving robust constraints on the

volumes of the cave and the main cone. Twelve locations from inside the cave at

Thríhnúkagígur were used to collect high resolution Light Detection and Ranging

(LiDAR) data to create a digital elevation model of the cave in 2011. The exterior of the

cone was also scanned. Data was collected with a Leica C-10 scanner. The resulting

DEM from the interior and exterior scans is accurate to <4 cm, yielding an accurate and

precise volume estimate of the main chamber of the cave and the spatter cone at the

surface.

35

2.3.2 Density Measurements

In order to compare the mass of erupted lavas and the mass of the buried tephra

that may have once filled the volume of the cave, the dense rock equivalent (DRE)

volumes of the two units must be determined. Vesicles in the lava samples were

determined to have a modal percentage of 29% by point counting. Most vesicles were at

least 1 mm in diameter; therefore lava chips were crushed with a mortar and pestle into

fragments that were <3 mm in any dimension and hand selected to ensure that all vesicles

were open. A Mettler electronic balance was used to weigh ~4 g of lava rock chips. A

similar mass of the better consolidated, yet still representative tephra were also weighed.

These samples were then placed in 10 ml and 50 ml graduated cylinders partially filled

with water, respectively, to measure the volume. Because the error on the 50 ml cylinder

was greater, a larger duplicate sample of the same tephra sample was also weighed and its

volume measured.

2.4 Results

The LiDAR derived DEM allows us to estimate that the volume of the cave is

~40,000 m3 (Figures 2.4, 2.12). The surface area of the lavas extruded from the

Thríhnúkagígur eruption can be estimated at 350,000 m2 from the geologic map by

Sæmundsson (2006). With a minimum estimated thickness of the lava flows, and cone

height and radius, a minimum erupted volume can be estimated. For this calculation, a 30

m high cone with a radius of 50 m and a lava flow thickness 0.25 m thick are used (Table

2.1). These estimates yield an erupted volume of 1.67·105 m3. The proportion of

36

incorporated tephra by mass that may have been entrained in the erupted lava can then be

determined using the DRE of the buried tephra and erupted lavas (Table 2.2). This

eruptive volume and a density of 2150 kg/m3 yields a total eruptive mass of 3.57 x 108 kg

of lava. The maximum possible mass fraction of tephra in the erupted Thríhnúkagígur

lavas is therefore 29%.

2.5 Discussion

A challenge for evaluating crustal assimilation in Iceland, and more generally at

ocean islands, is that these components have the same basaltic bulk compositions as the

magmas that intrude them. Without any knowledge of the physical and chemical

characteristics of potential crustal contaminants, it is nearly impossible to constrain their

contribution to mantle-derived magmas. The buried tephra could not have been identified

as a component in the Thríhnúkagígur lavas if only the lavas had been examined in

isolation. The unique exposure provided in the cave where the buried tephra is exposed

allows for it to be identified as a possible basaltic crustal contaminant. Direct field

observations and the physical properties of the crustal contaminant (fine-grained, poorly

consolidated, glassy, etc.) are critical to interpreting their role in magma production and

eruption dynamics.

2.5.1 Cave formation

Without a major edifice loading the crust at Thríhnúkar, and in the absence of a

central magma chamber, local asperities in the uppermost crust should be the secondary

37

control on the orientation of the dike after the regional stress field. Indeed, this is

observed in Thríhnúkagígur’s internal structure. The vertical sheet-like structure of the

dike along the length and height of the cave supports a dominant regional stress field

even in the uppermost crust, with the minimum stress in the north-northeast direction

(Clifton et al., 2002; Clifton and Kattenhorn, 2006; Keiding et al., 2008; Keiding et al.,

2009). The dike ruptured the surface up to 1.5 km away from Thríhnúkagígur to the

north-northeast where several smaller volume cones are mapped as part of the

Thríhnúkagígur event (Figure 2.1; Sæmundsson, 2006) demonstrating that it or a series of

en echelon dikes behaved uniformly over at least a couple of kilometers.

In addition to this primary control on dike orientation, the unconsolidated buried

tephra is much less competent than the overlying laterally continuous basalt flows, acting

as a less resistant pathway for magma to travel to the surface. This preexisting structure is

a local heterogeneity that likely guided and concentrated the magma during its final

ascent to the surface. Evidence of magmatic erosion or scouring of the tephra is preserved

in proto-conduits that penetrate into the buried tephra. These structures are seen in

contact with or immediately adjacent to the dike and are always seen intruding into the

tephra. The flowback-lined walls of proto-conduits in the deepest extent of the SW end of

the cave suggest that magma acted as an erosive fluid that entrained the fine-grained,

poorly consolidated buried tephra.

The location of the plugged conduit directly above the highest point in the

remnant buried tephra cone is consistent with the idea that at initial stages of eruption,

flow of magma could have been funneled by the tephra’s cone-like geometry to the

surface. Without a corresponding surface feature, the plugged conduit either did not

38

actually feed an eruption, or any cone that it may have constructed was subsequently

buried by the main cone of Thríhnúkagígur. Ultimately, the dike became more favorable

than the plugged conduit for erupting lavas at the surface. The location of the buried

tephra cone directly below the main cone may indicate that the tephra still acted to

concentrate the eruption at this location. This may also explain why the main cone of

three different monogenetic eruptions – the Thríhnúkar complex – all occur within a few

hundred meters of each other.

The eruption at Thríhnúkagígur was likely Hawaiian to mildly Strombolian in

style. Fire fountaining associated with the opening of the fissure would have focused into

the main cone over time. This would have also focused rising gas pockets, such as the 0.5

m long gas bubble observed in the dike deep at the SW end of the cave and may reflect a

shift to a more Strombolian eruption with punctuated explosions. Repeated, extremely

thin (2-5 cm) lava flows indicate flow back (Figure 2.3) into the vent and are consistent

with repeated bursts of lava erupting through the main conduit. Punctuated explosions

associated with slug flow would allow lava to drain back into the conduit repeatedly.

These observations point towards a low viscosity magma with intermediate to high gas

content, which is also consistent with the thin (~0.25 m), vesicle-rich lava flows and

Pele’s tears observed on the surface of the cone.

A plausible scenario to explain the formation and preservation of the cave beneath

Thríhnúkagígur is that unconsolidated or poorly consolidated material (the buried tephra)

was entrained by the erupting magma as the diking event progressed. Assuming that the

volume of the present day cave represents the maximum amount of tephra that could have

been incorporated into Thríhnúkagígur lavas, 10-29% of the lava could be derived from

39

the buried tephra cone. The buried tephra under Thríhnúkagígur, coupled with the

vacancy below the main cone, demonstrates that incorporation of poorly consolidated,

low-density material can occur rapidly, even during the course of an eruption if

conditions are favorable. Up to 29% by mass of tephra may be incorporated in this

manner.

The schematic diagram in Figure 2.13 illustrates a conceptual model for cave

formation in two views. The diagram on the left is a cross-section within the plane of the

dike that fed the eruption of Thríhnúkagígur. The right hand side is a diagram of a cross-

section perpendicular to the dike. The model begins with the eruption of a tephra cone

(Figure 2.13a) and then buried by subsequent lava flows (Figure 2.13b), which continue

to accumulate through time until 3500 BP. Figures 2.13c and 2.13d show the initial

stages of eruption where the dike cross-cuts the buried tephra cone and lava flows, and

then erupts at the surface. As the eruption progresses, tephra becomes entrained in the

ascending magma; and simultaneously the fissure transitions from erupting along its

entire length to focusing into a couple of vents (Figure 2.13e). At the final stage of

eruption, the dike has focused into a pair of adjacent vents that form the main spatter

cone (Figure 2.13f). Figure 2.13g shows the cave profile at present after cave-ins have

filled in part of the remnant buried tephra cone with debris.

2.6 Conclusions

Monogenetic fissure eruptions typically accommodate far field stresses, but the

orientation and propagation of dikes that intrude the crust or result in eruptions may also

40

be affected by a local stress field or preexisting crustal structures (Muirhead et al., 2015).

At Thríhnúkagígur, the orientation of the eruptive fissure was consistent with

accommodation of the far field stress regime, ascending to the surface as a vertical planar

dike. The preexisting buried tephra cone did not affect the geometry or orientation of the

dike as it propagated into the overlying lava flows to the surface; however, the location of

the main cone of Thríhnúkagígur probably resulted from the anisotropy produced by the

presence of a buried tephra cone at depth. The low density, fine-grained unconsolidated

tephra is less competent than the surrounding and lava flow that bury it, as demonstrated

by the numerous proto-conduits extending into it. The buried tephra very likely provided

a preferential pathway for the ascending magma thereby concentrating flow through

fissure at the main cone of Thríhnúkagígur. Here, evidence suggests that small-scale

structures in the shallow crust do not influence the orientation or geometries of fissure

eruption, but they may play a role in the location of the main vents that develop as the

eruption progresses.

Moreover, as the flow of magma was focused through the buried tephra, the

ascending magma scoured out, entrained, and erupted the unconsolidated buried tephra

with it. This the preferred model for the development of the subterranean cave below the

Thríhnúkagígur cone. Numerous proto-conduits fingers extend off the dike up into the

overlying tephra dike at depth in the cave, providing evidence for erosion of the buried

tephra by the ascending magma. A minimum estimate of the erupted volume of lava from

Thríhnúkagígur from field observations and a geologic map by Sæmundsson (2006) were

compared to the volume of the cave from LiDAR scans. The DRE of the buried tephra

41

and lava were then used to find that the maximum percent by mass of tephra that could be

in the lavas is 29%.

The rapid incorporation of the buried tephra into the erupting magma

demonstrates that magmas may be contaminated rapidly and to a significant extent by

shallow crustal components. If not for this unique exposure of the tephra in the cave, the

buried tephra could not have been identified as a component in the Thríhnúkagígur lavas

and this shallow crustal contamination would have been untraceable. The unconsolidated,

poorly crystalline, and fine-grained nature of the tephra facilitated its ingestion and

similar properties are advantageous for deeper crustal assimilation as well.

42

Table 2.1. Estimates of Thríhnúkagígur lavas volumes and calculation of tephra contents.

Cone

Height

(m)

Cone

Radius

(m)

Cone

Volume

(m3)

Flow

Area

(m2)

Flow

Thickness

(m)

Minimum

Erupted

Volume

(m3)

Dense

Rock

Equivalent

(kg/m3)

Minimum

Erupted

Mass (kg)

Cave

Volume

(m3)

Tephra

Density

(kg/m3)

Maximum

Tephra

Mass (kg)

% Tephra by

Mass

30 50 78,500 350,000 0.25 166,500 2150 3.57E+08 40,000 2550 1.02E+08 29%

Table 2.2. Densities of Thríhnúkagígur lava and buried tephras.

Sample ID Mass (g) Volume (mL) Density (kg/m3)

Lava TNG-14-33 4.268 1.5 2850

Buried tephra TNG-14-22 4.071 1.6 2540

Buried tephra TNG-14-32 13.801 5.4 2560

43

Figure 2.1. Map of the Thríhnukar fissure system modified from Sæmundsson (2006). Samples are shown in black stars. Subterranean samples are

excluded from this map, but are generally located underneath the Thríhnúkagígur spatter cone in red.

44

Figure 2.2. Two vents comprise the main cone of Thríhnúkagígur. The northern vent (left) became plugged towards the end of the eruption, while

the southern vent (right, with scaffolding) remained open, providing an entrance to the cave below. Photo from P. LaFemina.

45

Figure 2.3. Succession of thin (2-5 cm) lava flows back into the vent at the summit of Thríhnúkagígur. Photo from P. LaFemina.

46

Figure 2.4. LiDAR scan of the interior of the cave at Thríhnúkagígur. Green and blue coloring represents different reflectivities of the rock. The

dike, plugged conduit, and the buried tephra appear in blue. The image on the left is a view parallel to the orientation of the dike. The view on the

right is perpendicular to the orientation of the dike.

47

Figure 2.5. Magma flowback rivulets on a fallen block in the cave. Photo from P. LaFemina.

48

Figure 2.6. Stalactites of lava that dripped back into the cave during and/or following the eruption. Photo from P. LaFemina.

49

Figure 2.7. The NNE cave wall. A photo mosaic by P. LaFemina is simplified in a cartoon depicting the dike and tephra with sample locations.

50

Figure 2.8. Representative buried tephra (right; TNG-14-15) is poorly consolidated compared to the well-

welded hyaloclastite (left) collected at the from the gentle dipping hyaloclastite unit postulated by

Sæmundsson (2006).

51

Figure 2.9. A panorama of the dike at the north end of the cave. Here it cross-cuts the tephra apron and to the left interfingering between the buried

tephra and subsequent lava flows can be seen. Photo from P. LaFemina.

52

Figure 2.10. Photograph of the interior of the cave. The plugged conduit is the roundish purple feature on the left. The dike is the vertical purple-

black feature on the right. The buried tephra cone on the bottom left if black.

53

Figure 2.11. A ~0.5 m long gas pocket in the center of the dike.

54

Figure 2.12. LiDAR scan of the Thríhnúkagígur cone and cave with the location of the largest of the proto-conduits indicated (a). A LiDAR scan

of the main proto-conduit shows its geometry in greater detail (b).

55

Figure 2.13. Stages of Thríhnúkagígur formation starting with the formation (a) and burial of the tephra

cone (b), followed by dike injection (c) and eruption along a fissure (d). As the eruption continues, the

magma both progressively entrains more tephra from depth and narrows from a fissure into a couple of

vents (e anf f). Part g shows the present day outline of the cave. The left hand side of the diagram shows a

cross-section within the plane of the dike. The left shows a cross-section perpendicular to the dike.

56

Figure 2.13 (d-f) continued.

Figure 2.13 (g) continued.

References

Biggs J., Chivers M. and Hutchinson M. C. (2013) Surface deformation and stress

interactions during the 2007-2010 sequence of earthquake, dyke intrusion and

eruption in northern Tanzania. Geophys. J. Int. 195, 16–26.

Calais E., D’Oreye N., Albaric J., Delvaux D., Déverchère J., Ebinger C., Ferdinand R.

W., Kervyn F., Macheyeki A. S., Oyen A., Perrot J., Saria E., Smets B., Stamps D.

S. and Wauthier C. (2008) Strain accommodation by slow slip and dyking in a

youthful continental rift, East Africa. Nature 456, 783–788.

Clifton A. E. and Kattenhorn S. A. (2006) Structural architecture of a highly oblique

divergent plate boundary segment. Tectonophysics 419, 27–40.

Clifton A. E., Sigmundsson F., Feigl K. L., Gunnar Guðmundsson and Árnadóttir T.

(2002) Surface effects of faulting and deformation resulting from magma

accumulation at the Hengill triple junction, SW Iceland, 1994-1998. J. Volcanol.

Geotherm. Res. 115, 233–255.

Fronval T. and Jansen E. (1997) Eemian and Early Weichselian (140–60 ka)

Paleoceanography and paleoclimate in the Nordic Seas with comparisons to

Holocene conditions. Paleoceanography 12, 443.

Genareau K., Valentine G. A., Moore G. M. and Hervig R. L. (2010) Mechanisms for

transition in eruptive style at a monogenetic scoria cone revealed by microtextural

analyses (Lathrop Wells volcano, Nevada, U.S.A.). Bull. Volcanol. 72, 593–607.

Harp A. G. and Valentine G. A. (2015) Shallow plumbing and eruptive processes of a

scoria cone built on steep terrain. J. Volcanol. Geotherm. Res. 294, 37–55.

Keating G. N., Valentine G. A., Krier D. J. and Perry F. V. (2007) Shallow plumbing

systems for small-volume basaltic volcanoes. Bull. Volcanol. 70, 563–582.

Keiding M., Árnadóttir T., Sturkell E., Geirsson H. and Lund B. (2008) Strain

accumulation along an oblique plate boundary: The Reykjanes Peninsula, southwest

Iceland. Geophys. J. Int. 172, 861–872.

Keiding M., Lund B. and Árnadóttir T. (2009) Earthquakes, stress, and strain along an

obliquely divergent plate boundary: Reykjanes Peninsula, southwest Iceland. J.

Geophys. Res. Solid Earth 114, 1–16.

59

Muirhead J. D., Kattenhorn S. A. and Le Corvec N. (2015) Varying styles of magmatic

strain accommodation across the East African Rift. Geochemistry Geophys.

Geosystems, 1541–1576.

Németh K. and Kereszturi G. (2015) Monogenetic volcanism : personal views and

discussion. Int. J. Earth Sci. 104, 2131–2146.

Németh K. and White J. D. L. (2003) Reconstructing eruption processes of a Miocene

monogenetic volcanic field from vent remnants: Waipiata Volcanic Field, South

Island, New Zealand. J. Volcanol. Geotherm. Res. 124, 1–21.

Sæmundsson K. (2008) Þríhnúkagígur Helstu niðurstöður kjarnaborana.

Sæmundsson K. (2006) Þríhnúkagígur Jarðfræðirannsóknir og tillögur.,

Sæmundsson K., Jóhannesson H., Hjartarson Á., Kristinsson S. G. and Sigurgeirsson M.

Á. (2010) Geological Map of Southwest Iceland, 1:100,000.

Sigmundsson F., Hooper A., Hreinsdóttir S., Vogfjörd K. S., Ófeigsson B. G., Heimisson

E. R., Dumont S., Parks M., Spaans K., Gudmundsson G. B., Drouin V., Árnadóttir

T., Jónsdóttir K., Gudmundsson M. T., Högnadóttir T., Fridriksdóttir H. M., Hensch

M., Einarsson P., Magnússon E., Samsonov S., Brandsdóttir B., White R. S.,

Ágústsdóttir T., Greenfield T., Green R. G., Hjartardóttir Á. R., Pedersen R.,

Bennett R. a, Geirsson H., La Femina P. C., Björnsson H., Pálsson F., Sturkell E.,

Bean C. J., Möllhoff M., Braiden A. K. and Eibl E. P. S. (2015) Segmented lateral

dyke growth in a rifting event at Bárðarbunga volcanic system, Iceland. Nature 517,

191–5.

Stefánsson Á. B. (1992) Þríhnúkagígur. Náttúrufræðingurinn 61, 229–242.

Valentine G. A. (2012) Shallow plumbing systems for small-volume basaltic volcanoes,

2: Evidence from crustal xenoliths at scoria cones and maars. J. Volcanol.

Geotherm. Res. 223-224, 47–63.

Wood C. A. (1980) Morphometric evolution of cinder cones. J. Volcanol. Geotherm. Res.

7, 387–413.

Wright T. J., Ebinger C., Biggs J., Ayele A., Yirgu G., Keir D. and Stork A. (2006)

Magma-maintained rift segmentation at continental rupture in the 2005 Afar dyking

episode. Nature 442, 291–294.

60

Chapter 3

Magma genesis at Thríhnúkagígur, Reykjanes Peninsula, Iceland:

Implications for magmatic plumbing structure at monogenetic volcanic fields

3.1 Introduction

The Reykjanes Peninsula, in the absence of central volcanoes and large

homogenizing magma chambers, is dominated by monogenetic basaltic fissure eruptions.

With a few exceptions (Gee et al., 1998a; Gee et al., 2000), the geochemistry of lavas on

the Reykjanes Peninsula has been interpreted to primarily reflect the mantle sources that

generated these melts. Recent unrest in the region (Clifton et al., 2002; Michalczewska et

al., 2012) reflects that magma is injected and stored in the crust, so it is necessary to

understand both the magmatic plumbing structure of these intrusions and the effect of

magma storage in has on the geochemical evolution of these magmas.

Radiogenic isotope analysis has traditionally formed the bulk of the mantle source

characterization for Reykjanes Peninsula lavas, and lavas across Iceland more generally

because of the plume-ridge interactions and intrinsic chemical heterogeneities within the

plume. However, some authors suggest that these isotopes, and Sr in particular, could be

affected by interaction with or assimilation of hydrothermally altered basalts during a

magma’s residence in the crust (Gee et al., 1998a; Gee et al., 2000). Because mantle

sources have been debated extensively with samples from a much greater geographic area

than the single location of Thríhnúkagígur (Condomines et al., 1983; Hanan and

Schilling, 1997; Chauvel and Hémond, 2000; Skovgaard et al., 2001; Stracke et al., 2003;

61

Kokfelt et al., 2006; Kitagawa et al., 2008; Peate et al., 2010; Koornneef et al., 2012;

Rudge et al., 2013; Shorttle et al., 2013; Manning and Thirlwall, 2014), source

geochemistry is not the main focus of this study. Rather, results of Sr-Nd-Pb isotope

ratios in lavas from Thríhnúkagígur are discussed primarily within the context of the

literature that has attempted to define specific mantle source compositions on the

Reykjanes Peninsula (Hanan and Schilling, 1997; Thirlwall et al., 2004; Kokfelt et al.,

2006; Kitagawa et al., 2008), to evaluate the potential to use Sr as a tracer of crustal

contamination of Thríhnúkagígur magmas.

Using a combined major and trace element and Sr-Nd-Pb isotope geochemical

approach this work seeks to understand, from source to surface, the evolution of magma

erupted at Thríhnúkagígur. The unique exposure offered by the cave below the main cone

of Thríhnúkagígur allows sampling of the intrusive part of the system and older buried

tephra in addition to the surficial lavas. Heterogeneity in the compositions from different

part of the system suggests a magmatic plumbing structure that does not homogenize

mantle-derived melts. Geophysical and geodetic observations of other intrusions and

fissure eruptions provide a context in which to invoke a magmatic plumbing system that

may preserve geochemical variations.

3.2 Sample descriptions

At the surface, one Miðhnúkur and four Thríhnúkagígur lava samples were

collected for analysis (Figure 2.1). A sample of vent spatter at the summit of

Thríhnúkagígur from a previous sampling campaign was also prepared for analysis.

Within the cave, 5 dike samples, one sample of the fallen intrusive plug, and 6 buried

62

tephra samples were collected for geochemical analysis. Photograph mosaics and

sketches provide sample locations within the main chamber of the cave (Figures 2.4, 2.5).

3.2.1 Surface lavas

Lavas from Thríhnúkagígur are highly vesicular (18-36.5% by volume) and

relatively glassy with <15% plagioclase phenocrysts (Table 3.1; Figure 3.1a, b). Samples

TNG-14-33 and TNG-14-34 were collected from some of the thicker flow tops (~20 cm)

in the area as outcrops with exposures of the interior of the H146 flow were not found.

TNG-14-37 was taken from the lip of a former lava lake. TNG-14-32 is from a small

flow from a flank vent on the main cone that traveled into the lava lake depression after

the lava had cooled and/or drained.

One sample, TNG-14-35, was collected from a laterally more extensive, older

Holocene flow from Miðhnúkar (H145). Miðhnúkar flows are easily distinguishable from

the Thríhnúkagígur lavas as they contain macroscopic olivine phenocrysts (Figure 3.1f)

rather than plagioclase.

3.2.2 Dikes

The dike is heterogeneous and samples were taken from a variety of different

outcrops. Samples were taken from the interior of the dike, a 6 cm thin dike that splays

off of the main dike, and a chilled margin of the dike. Samples from the interior of the

dike tend to be more vesicular, have larger plagioclase phenocrysts, and courser

groundmass compared to the margin of the dike. These changes are typically gradational,

63

but TNG-14-27e (exterior; Figure 3.1c) and TNG-14-27i (interior; Figure 3.1d) shared a

sharp contact. These two samples are the most contrasting dike samples as TNG-14-27i

has flow banding and interstitial clinopyroxene between the plagioclase phenocrysts,

while TNG-14-27e is devoid of phenocrysts and vesicles, and has a fine-grained

groundmass.

3.2.3 Plugged conduit

This is the only sample that was not collected in situ. TNG-14-18 was collected

immediately below an intrusive, partially hollow structure on the cave ceiling.

Weathering of the cave walls to a yellow color contrasted the reddish-purple altered

surfaces of this feature. It is notable for being aphanitic with very few phenocrysts.

3.2.4 Buried tephras

Tephra samples consist of primarily glassy, vesicular clasts (Figure 2.8); however,

the total crystallinity, gas content, grain size distribution, and degree of alteration of the

clasts is variable. Samples may contain clasts that are exclusively fresh and black or have

a mixture of black and brown glasses. Clasts that are larger than 5 mm are present in most

samples although they rarely exceed 2-3 cm. Grain mounts were made for 4 of the 6

tephra samples. The grain mounts excluded clasts larger than 5 mm. The two exceptions

for which no grain mounts were made are TNG-14-17, which is a fist sized scoria clast

and TNG-14-28, which was consolidated and contained some sedimentary structure.

Neither is representative of the bulk tephra pile. Free whole crystals of plagioclase and

64

olivine occur, while broken fragments of large plagioclase crystals are the most common

crystalline clast type. There are also rare plagioclase and plagioclase-olivine

glomerocrysts.

3.3 Methods

Samples were cut into centimeter thick slabs and the weathered edges were

trimmed off of the slabs with a ceramic tile saw and polished to remove saw marks. Thin

section billets were cut from the slabs and sent to Spectrum Petrographics to make

standard 30 μm microprobe polished thin sections. The remaining slabs for each sample

were chipped into millimeter sized pieces with a ceramic jaw crusher. Tephras were

rinsed in an ultrasonic bath of milliQ water for 1 hour and dried. Tephras and rock chips

were transferred into a tungsten carbide shatterbox and run for 20-30 seconds to create a

fine powder.

3.3.1 Point counting

Mineral modes were determined by point counting of thin sections under a

petrographic microscope. Grids of 10 by 20 were used to achieve 200 total counts per

thin section. For the grain mounts of the buried tephra, point counting continued until 150

spots on grains had been counted. For the other thin sections, vesicles were included in

the total of 200 counts.

65

3.3.2 Major element analytical methods

Bulk rock major and trace element compositions were determined at the

Laboratory for Isotopes and Metals in the Environment at The Pennsylvania State

University. Lithium metaborate fusions were made and prepared in nitric acid for major

element analysis. Splits of 100 mg sample powder were weighed and mixed with 1.000 g

of lithium metaborate in preweighed vials. After gently shaking the mixture, it was

transferred to graphite crucibles and placed in a furnace preheated to 900°C. The

crucibles were heated to 1000°C for 10 minutes, then cooled back to 900°C. The

resulting melts were added to 100 ml of 5% HNO3 solutions in Teflon vials with

magnetic stir bars. The solutions were stirred for at least 15 minutes before being

transferred to polyethylene containers for storage. These solutions were analyzed for

major element concentrations using a Perkin-Elmer Optima 5300DV Inductively Coupled

Plasma Atomic Emission Spectrometer (ICP-AES).

3.3.3 Trace element analytical methods

For trace element analysis, 100 mg of powder were weighed out for each sample.

Powders were subsequently digested in three steps; first with a combination of 6N HF

and 8N HNO3, then with 2N HCl, and finally with 4N HNO3. Samples were dried down

between each stage of digestion. Following the final dry down, samples were taken up in

HNO3 and diluted with milliQ water to 2000 times their original weight in a matrix of

~2% HNO3. High concentrations of Ba and Sr required a separate dilution to 5000 times

the original sample weight, also in a ~2% HNO3 solution in milli-Q water. This

66

procedure was also followed for a method blank and five rock standards: BHVO-1, BCR-

1, BR, BIR-1, and JA-1.

Trace elements were analyzed using a Thermo X-Series II Quadrupole

Inductively Coupled Plasma Mass Spectrometer (ICP-MS). The instrumental background

was determined using a 2% HNO3 solution in milli-Q water. Background was found to be

≤0.01 for all elements except Ni and Sc, which were less than 0.1 ppb, and Cu, Cr, and

Zn, which were less than 0.54 ppb. The experimental design started with a method blank

and the five rock standards. The method blank was run again before the samples were

analyzed as unknowns. Calibration lines with an R2 value of greater than or equal to

0.999 were used for determining concentrations of elements in the unknowns. One

standard data point was removed to produce curves for several transition metals and rare

earth elements. Two standard data points had to be removed from the Co, and Ho curves

to produce an R2 of 0.999. The true dilution factors for each sample were determined

based on the initial mass of the powdered sample and the mass of the final diluted sample

for analysis. The concentrations in solution were then multiplied by the dilution factor

(~2000 except for ~5000 for Ba and Sr) to determine the concentration of each analyte in

the bulk rock.

Samples were digested and analyzed in two separate batches. Samples TNG-14-

18 and TNG-14-27e were duplicated in the second analysis. For these replicate analyses,

reported values for most analytes are within 5.0%. The elements Sc, V, Co, Cu, Sr, Ce,

Lu, and U were all within 8.0%. Only Cs (39.0%) and Ta (18.3%) had larger errors.

These two elements occur in concentrations of < 0.5 ppm in these samples and are not

used to predicate any arguments presented herein.

67

3.3.4 Pb, Sr, and Nd isotope analytical methods

Acid digestion, ion exchange chromatography, and isotope ratio measurements of

Pb, Sr, and Nd were conducted at the Consiglio Nazionale delle Ricerche (CNR), Pisa,

Italy. Splits of 200 mg were weighed out for each sample and dissolved in a combined

concentrated HF and HNO3 solution, then HNO3, followed by 7N HCl, and finally with

0.7N HBr. Sample solutions were dried down between each step. Samples were

transferred to centrifuge tubes and centrifuged to remove any residuum before ion

exchange chromatography. Acid digestion and ion exchange chromatography procedures

are described in Appendix A.

Isotope ratios were measured on a Finnigan MAT 262 multi-collector thermal

ionization mass spectrometer (TIMS) running in static mode for Pb and dynamic mode

for Sr and Nd. For Pb isotopic analysis, mass discrimination corrections of 0.15 ± 0.01%

per mass unit relative to the NIST SRM 981 reference standard were applied to

experimental data (Todt et al., 1993). Following this correction, Pb isotope analyses have

2SD within 0.025% per mass unit. Laboratory Pb blanks are routinely between 0.2-0.4

ng, and no blank correction is made.

Ratios of 87Sr/86Sr are normalized to 86Sr/88Sr of 0.1194 and the mean 143Nd/144Nd

ratios are normalized to 146Nd/144Nd of 0.7219. The results of 19 analyses of Sr standard

NIST SRM 987 (SrCO3) gave an average of 0.710227 ± 20 (2SD). Measured values are

corrected to 87Sr/86Sr = 0.710250. Blanks for Sr are <0.3 ng and are negligible. Twenty-

seven analyses of the Nd standard is JNdi-1 yield 0.512104 ± 10 (2SD) and no correction

was made.

68

3.4 Results

3.4.1 Major elements

Lavas from the two eruptions in Thríhnúkagígur fissure system and the buried

tephras exposed at depth within the cave are basalts (Table 3.2; Figure 3.2; Le Bas and Le

Maitre, 1986) with 46.70 – 49.25 wt.% SiO2 and 6.03 – 9.58 wt.% MgO (Mg# 47 – 61).

This range is broadly consistent with tholeiitic Holocene lavas of the Reykjanes

Peninsula (Jakobsson et al., 1978; Gee et al., 1998a; Peate et al., 2009). The dike from the

most recent eruption has 47.23 – 49.25 wt.% SiO2 and 10.78 – 13.20 wt.% Fe2O3T at

relatively constant MgO content (7.54-8.04 wt.%). The buried tephra is the only rock unit

with samples that have a non-zero LOI (0.69 – 2.79 wt.%), suggestive of variable

alteration or weathering. These samples have 14.42 – 16.09 wt.% Al2O3 and 46.70 –

48.38 wt.% SiO2. By comparison, the lavas fed by the main dike are more homogeneous

(Al2O3 = 15.2-15.8 wt.%; SiO2 = 47.7-48.2 wt.%). All unaltered samples have CIPW

norms that are olivine normative (Table 3.3).

Plots of major elements versus MgO show a compositional gap between the

Holocene basalts (MgO ≥ 7.25 wt.%) and the older buried tephras (MgO ≤ 6.78 wt.%).

Thríhnúkagígur lavas and dikes have a steep, positive linear trend for Al2O3 vs MgO,

differing from the buried tephra, which have relatively constant Al2O3. For CaO, the lava

and dike samples show limited compositional range (12.00-13.38 wt.%), while the buried

tephras have lower concentrations (9.79-11.43) that give the whole suite a weakly

positive trend. TiO2 and Fe2O3 increase linearly with decreasing MgO, although several

samples depart from the overall trend (Figure 3.3). The chilled margin (TNG-14-27e) has

69

low TiO2 and Fe2O3 for its MgO while the plugged conduit (TNG-14-18) and Miðhnúkar

(TNG-14-35) have high TiO2 and Fe2O3 for their MgO contents. Concentrations of SiO2

increase with decreasing MgO, but this trend is not continuous across the entire sample

suite (Figure 3.3). Rather, the lava flows and dike samples form a separate trend that is

offset to higher silica for their MgO content relative to the buried tephra samples. One

dike sample from the chilled margin (TNG-14-27e) markedly departs from that slope

with ~1 wt.% more SiO2 than the other lava and dike samples. The alkalis form linear

negative trends with MgO.

The chemical index of alteration (CIA; Nesbitt and Young, 1982) was used to

evaluate the extent of alteration (Figure 3.3). The CIA is defined as 100*[Al2O3/(Al2O3 +

CaO + Na2O + K2O)]. Two visibly altered samples – a surface vent spatter sample (MC-

03-612) and a reworked, partially palagonitized tephra (TNG-14-28) – had a CIA of 58,

compared to CIAs of ≤53 in all other samples. These samples depart from otherwise

linear trends in several major elements when plotted against MgO (Figure 3.3). Values of

CIA for fresh basalts are typically 30-45 (Nesbitt and Young, 1982), so the samples in

this study may have all experienced mild alteration or may have incorporated altered

crust during their ascent through the crust.

3.4.2 Trace elements

Concentrations of the compatible trace elements Ni and Cr decrease with

decreasing MgO. Scandium varies by rock type; buried tephras show a positive trend

with MgO while the Holocene lavas and dikes show a negative correlation.

70

Concentrations of the incompatible elements Sr, Eu, and Pb are inversely correlated with

MgO (Figure 3.4; Table 3.4).

Primitive mantle normalized multi-element diagrams are characterized by strong

positive Nb and Ta anomalies, a positive Ba anomaly, and a large negative Pb anomaly

for all samples (Figure 3.5; Sun and McDonough, 1989). The vent spatter (MC-612-03

with CIA = 58) is the sole exception with a positive Pb anomaly, which is interpreted to

be the result of contamination. The average values for each of the rock units yield similar

patterns for most incompatible trace elements. The buried tephras are enriched relative to

the lava and dike suite, but form generally parallel trends. Strontium has a negative

anomaly in the buried tephras and a positive anomaly for the chilled margin (TNG-14-

27e) and the plugged conduit (TNG-14-18).

Chondrite normalized rare earth element (REE) diagrams display relatively

smooth patterns, decreasing steadily from the light REEs to the heavy REEs (Figure 3.6;

Sun and McDonough, 1989). The chilled margin (TNG-14-27e) and plugged conduit

(TNG-14-18) have the flattest REE patterns ([La/Yb]N = 1.92 and 1.68, respectively) in

samples from Thríhnúkagígur. The buried tephras are the most enriched in all REEs, but

are relatively more enriched in the LREEs, giving rise to a steeper pattern ([La/Yb]N >

2.75). The lavas and the dike samples are intermediate between the buried tephras and the

plugged conduit in terms of REE composition. There is a very subtle positive Eu anomaly

in the chilled margin of the dike (TNG-14-27e). However, no other samples from

Thríhnúkagígur have Eu anomalies, positive or negative.

71

3.4.3 Pb, Sr, and Nd isotopes

There is a positive relationship in the 206Pb/204Pb – 208Pb/204Pb data (Figure 3.7a).

The most radiogenic samples are the buried tephras, although one Holocene

Thríhnúkagígur sample has higher 208Pb/204Pb (Table 3.5). The least radiogenic sample is

the vent spatter (MC-612-03). This sample has exceptionally high Pb concentration (2.40

ppm) and a CIA of 58. It has a large positive Pb anomaly on the primitive mantle

normalized diagram, while all other samples have strong negative Pb anomalies (Figure

3.5). The vent spatter has very similar major and trace element abundances to other lavas

and dikes. This suggests that Pb contamination was added to the sample, so it will not be

considered as representative of any trend or end member. The most unradiogenic sample

considered to be primary is from the dike (TNG-14-24). In general, the Holocene lavas

and dikes have a greater range of 206Pb/204Pb and 208Pb/204Pb than the buried tephra,

especially to lower 206Pb/204Pb ratios.

Ten of eighteen samples plot in a relatively narrow range in 206Pb/204Pb and

207Pb/204Pb (Figure 3.7b). In these 10 samples, ratios of 206Pb/204Pb are between 18.79 and

18.95 and 207Pb/204Pb ratios range from 15.47 and 15.55 (Table 3.5). All but one of the

buried tephras fall within this range. Two additional dike samples fall within this range of

206Pb/204Pb, but with higher 207Pb/204Pb ratios. The remaining six samples form a negative

linear trend (Figure 3.7b).

Ratios of 143Nd/144Nd range from 0.512982 to 0.513066. Despite the narrow

range, the buried tephras are distinct from the lavas and dike samples. With the exception

of the most altered buried tephra sample (TNG-14-28 with CIA = 58), the buried tephras

do not exceed 0.513018 ± 0.000010. A single lava sample has a 143Nd/144Nd value within

72

the tephra range, while all other dike and lava samples have ratios greater than 0.513021.

There is some slight overlap between the buried tephra suite and the dike and lava

samples when error is considered (Table 3.5). The plugged conduit (TNG-14-18) has a

ratio of 0.513039 ± 0.000008, placing it at the high end of the measured 143Nd/144Nd

values.

Three dike samples define the lowest 87Sr/86Sr values of 0.703135 to 0.703143

(Figure 3.8). The other lavas and dikes span the same range of Sr isotope ratios as the

buried tephra. Aside from the buried tephra with the high CIA, only a single lava sample

falls within the reported values for the buried tephras in Nd ratios. The most depleted

(unradiogenic Sr and radiogenic Nd) sample is the same as the most depleted Pb sample,

dike sample TNG-14-24.

3.5 Discussion

3.5.1 Source geochemistry

3.5.1.1 Trace element characteristics

High field strength (HSFE) and rare earth element (REE) ratios illustrate

important differences between SW Iceland (RP and WVZ) and the Eastern Volcanic

Zone. The Reykjanes Peninsula and the Western Volcanic Zone comprise a steeper

positive trend separate from the EVZ on a plot of (La/Yb) N vs. Nb/Zr (Figure 3.9),

suggesting a uniquely Nb-enriched source component in the west. Thríhnúkagígur buried

tephras have the highest (La/Yb)N among the RP and some of the highest Nb/Zr in all of

73

Iceland. Their Nb/Zr ratios are comparable to those of the most enriched EVZ volcanics,

but at much lower (La/Yb)N. If similar mantle source geochemistry is assumed, then

incompatible trace element ratios differences may attributed to melting in the presence of

garnet. The thicker lithosphere in the EVZ should generate deeper mantle melts than on

the RP, so the difference in the SW Iceland and EVZ slopes on the (La/Yb) N vs. Nb/Zr

plot could possibly be attributed to the relative compatibility of Yb garnet in southern

EVZ source region. This is consistent with the slow spreading rates associated with

incipient rifting in the southern EVZ as it propagates to the south (LaFemina et al., 2005).

Ratios of Nb/U, like Nb/Zr, suggest that Nb is anomalously enriched in the RP

mantle relative to the rest of Iceland (Kokfelt et al., 2006; Koornneef et al., 2012),

pointing to different geochemical source domains across the southern Iceland mantle.

Globally, magmas produced at ocean islands and mid-ocean ridges have Nb/U ratios of

47 ± 10 irrespective of Nb bulk concentrations, yet lavas from SW Iceland depart

markedly from this range, with Nb/U as high as 100 (Figure 3.10a; Hofmann et al.,

1986).

A similar pattern is observed for Ce on the RP. Ratios of Ce/Pb remain

remarkably constant for OIBs (25 ± 5; Hofmann et al., 1986), but the Reykjanes

Peninsula and the WVZ are more enriched with values commonly up to 40 (Figure 3.10b;

Hofmann et al., 1986). Rare samples have even higher Ce/Pb ratios up to 51 in the

Western Volcanic Zone and on the RP (Koornneef et al., 2012), including two

Thríhnúkagígur samples.

Strong enrichments in Nb/U and Ce/Pb ratios cannot be attributed to the melting

process because both of these ratios use elements that are nearly equally incompatible.

74

Unlike Yb, which is preferentially compatible in garnet, the source mineralogy does not

exert a differentiating effect on Nb relative to U, or Ce relative to Pb. This suggests that

there is a mantle component unique to the RP that contributes strongly to Thríhnúkagígur

magmas, and the RP more broadly.

3.5.1.2 Sr-Nd-Pb isotope systematics: mantle sources

The melt fraction, melting depth, and source mineralogy can strongly affect trace

element ratios, but do not affect isotope ratios. Several studies have proposed Sr-Nd-Pb

isotopic compositions for mantle sources in Iceland (Hanan and Schilling, 1997;

Thirlwall et al., 2004; Kokfelt et al., 2006; Kitagawa et al., 2008). They converge on

three main sources – an enriched plume source, an EM1-like source, and a depleted

(MORB-like) source. The EM1-like source is specific to the Reykjanes Peninsula in

Holocene basalts (Thirlwall et al., 2004), but is recorded in Tertiary basalts in eastern

Iceland (Hanan and Schilling, 1997; Kitagawa et al., 2008).

Lavas from south Iceland plot primarily along an array between the enriched

plume source and the depleted source in Sr-Nd-Pb isotope space and are subparallel to

the Northern Hemisphere Reference Line on Pb-Pb plots (NHRL; Hart, 1984). The

Reykjanes, Krísuvík, and Brennisteinsfjöll fissure swarms, all part of the Reykjanes

Peninsula, define largely overlapping fields on the 206Pb/204Pb vs. 208Pb/204Pb plot (Figure

3.11; Thirlwall et al., 2004; Kokfelt et al., 2006; Peate et al., 2009). These three fissure

swarms are bracketed by the less radiogenic Pb composition of southern Reykjanes Ridge

(Thirlwall et al., 2004) and the more radiogenic Pb compositions of the volcanic systems

of Hekla (Park 1990; Furman et al., 1995) and Vestmannaeyjar (Furman et al., 1991;

75

Chauvel & Hémond, 2000; Kokfelt et al., 2006; Peate et al., 2010) in the southern EVZ.

Katla has the most radiogenic Pb isotope ratios in south Iceland (Park 1990; Furman et

al., 1995) and acts as the enriched regional end member for both Pb-Pb plots. However,

some Reykjanes Peninsula lavas with lower 206Pb/204Pb are pulled off the NHRL to

higher 208Pb/204Pb relative to 206Pb/204Pb, towards the EM1-like component.

Thríhnúkagígur lavas and dikes plot with other Reykjanes Peninsula lavas on the

206Pb/204Pb vs. 208Pb/204Pb plot (Figure 3.11). The buried tephras and the majority of

Thríhnúkagígur lavas and dikes fall on the binary mixing array defined by south Iceland

lavas between an enriched plume source and a depleted MORB-like source, although

several Thríhnúkagígur lava and dike samples (MC-612-03, TNG-14-24, TNG-14-34,

TNG-14-18, and TNG-14-17) extend slightly off of the array and off of the NHRL

towards the EM1-like source.

The lowest 206Pb/204Pb sample (MC-612-03) is vent spatter from the summit of

Thríhnúkagígur, however, it is not shown on figures in this section because it is

considered to be contaminated (see section 3.4.3). Thríhnúkagígur lavas and dikes define

a field at 206Pb/204Pb ratios consistent with previously reported Brennisteinsfjöll

compositions. Excluding the vent spatter, the field defined by this single eruption at

Thríhnúkagígur is as broad as the entire reported range of previously reported data from

Brennisteinsfjöll (n = 9). Contributions from the EM1-like component are consistent with

some the Thríhnúkagígur lavas and dikes and other Reykjanes Peninsula lavas being

pulled off the NHRL. The buried tephra on the other hand, does not require a contribution

from this source.

76

While only a single enriched plume sources is revealed by the Pb-Pb isotope plot,

an enriched, high-Nb component unique to SW Iceland was identified on the basis of

Nb/U ratios. Plotted against 206Pb/204Pb (Figure 3.12a) and 208Pb/204Pb (Figure 3.12b),

there is a positive correlation between Nb/U ratios and Pb isotope ratios on the

Reykjanes, suggesting that there is a second enriched plume source in south Iceland with

the same Pb isotope compositions but with enriched Nb.

As with the Pb-Pb plots, Figure 3.13 shows that 87Sr/86Sr vs. 143Nd/144Nd ratios

from south Iceland form an array that is bracketed by Katla (relatively radiogenic Sr and

unradiogenic Nd; Park, 1990; Furman et al., 1995) on one end and the southern

Reykjanes Ridge (unradiogenic Sr and radiogenic Nd; Thirlwall et al., 2004) on the other.

Again, the enriched plume source and depleted MORB-like source bound the array,

which includes the Reykjanes Peninsula fissure swarms. Aside from a single Krísuvík

sample that plots with the EM1-like source at higher 87Sr/86Sr, there is no systematic

trend towards this mantle component in the RP data.

Thríhnúkagígur lavas and dikes with the lowest 87Sr/86Sr ratios cluster closely

with previously reported compositions at Brennisteinsfjöll on a plot of 87Sr/86Sr vs.

143Nd/144Nd (Figure 3.13). Most of the other lava and dike samples fit more broadly with

other Reykjanes and Krísuvík lavas, but fall outside the previously defined range of

Brennisteinsfjöll Sr-Nd isotope compositions (Thirlwall et al., 2004; Kokfelt et al., 2006;

Peate et al., 2009). One lava and one buried tephra fall outside of the previously

recognized Reykjanes Peninsula fields, extending in the direction of the EM1-like

component.

77

The Thríhnúkagígur data overlap on Sr-Nd-Pb isotope plots not only with other

Reykjanes Peninsula fissure swarms, but also with lavas from Vestmannaeyjar. The

relatively enriched buried tephra field overlaps with the depleted end of the

Vestmannaeyjar field on the 206Pb/204Pb vs. 208Pb/204Pb plot (Figure 3.11), and the

Thríhnúkagígur lavas, dikes, and buried tephras all fall within the Vestmannaeyjar field

in Sr-Nd isotope space (Figure 3.13). This affinity with Vestmannaeyjar suggests that

similar mantle components may contribute to magma genesis across south Iceland.

Vestmannaeyjar may be the mostly likely location to express geochemical similarities to

the Reykjanes Peninsula because neither the RP nor Vestmannaeyjar have central

volcanoes or are thought to have large magma chambers (unlike Hekla or Katla).

Therefore, their magmas are likely to experience similar processes and evolve by similar

pathways. For example, injections of magma into the crust are have short residence times

in both locations, on the order of decades, and feed small volume (<1 km3) monogenetic

fissure eruptions (Sigmarsson, 1996; Higgins and Roberge, 2007; Koerberle et al., 2015).

Assuming similar isotopic sources, variations between these volcanic systems

may be ascribed to the relative proportions of the different mantle sources contributing to

melt generation, which may be affected by differences in melting depth and melt fraction.

(While this explains isotopic similarities, it still fails to reconcile the differences in Nb/U

between the RP and Vestmannaeyjar as expressed in Figure 3.9, requiring an additional

mantle source.) Additionally, continuous melt extraction at the neovolcanic zones like the

Reykjanes Peninsula could gradually deplete enriched mantle sources. As an enriched

mantle component is repeatedly melted and becomes increasingly less fertile, its

proportional contribution relative to the depleted component is also likely to decrease.

78

The buried tephra is conservatively estimated to be at least 100 ka in age and likely older

because it is believed to have formed during the previous interglacial, which ended at 113

ka (Fronval and Jansen, 1997). Therefore, one possible explanation for the subtle

difference in the Pb and Nd isotopes in the buried tephra and the Holocene lavas and

dikes may be their age difference and the depletion of the enriched plume source beneath

the Reykjanes Peninsula through time.

In addition to the enriched, depleted, and EM1-like mantle sources already

discussed, Thirlwall et al. (2004) propose a fourth mantle source that falls along the south

Iceland array of lava compositions on Pb-Pb and Sr-Nd isotope plots. The contributions

of this source are unclear. This component would have a more moderate 87Sr/86Sr ratio

(~0.7032) than the EM1-like component, and may explain the trend of Thríhnúkagígur

samples on a Sr-Pb isotope plot (Figure 3.14). As with previous isotope plots, the

southern Reykjanes ridge is the most depleted of the south Iceland volcanic systems and

Katla is the regionally enriched end member. The Thríhnúkagígur data fall within the

enriched end of the Reykjanes field, beyond the Brennisteinsfjöll field, with two samples

recording higher 87Sr/86Sr ratios than previously reported for the Reykjanes Peninsula at

206Pb/204Pb ≥ 18.4. While this proposed mantle end member appears promising in Sr-Pb

isotope space, it cannot explain the high 87Sr/86Sr ratios of the sample suite at more

radiogenic Nd ratios. This suggests that the previously defined fourth end member of

Thirlwall et al. (2004) is not the main source of the Sr isotope signature in

Thríhnúkagígur lavas, dikes, or buried tephra, but neither can it be ruled out as a potential

contributor. Another source is still necessary to explain the radiogenic Sr isotope

signature.

79

3.5.1.3 Sr isotopes: potential crustal contributions

The low 206Pb/204Pb lava and dike samples that may require an EM1-like

component are not the same samples that trend toward the EM1-like source in Sr-Nd

isotope space. Dike sample TNG-14-24 has lowest 206Pb/204Pb (excluding vent spatter,

which is interpreted to be contaminated) and lowest 87Sr/86Sr and highest 143Nd/144Nd of

the lava and dikes. If TNG-14-24 reflected a larger contributions from an EM1-like

source, it would be expected to have the highest or among the higher 87Sr/86Sr samples

measured. Likewise, the buried tephra extend more radiogenic Sr ratios (Figure 3.13), but

do not extend towards the EM1-like component on Pb-Pb plots (Figure 3.11). The

inconsistent behavior of individual samples and the buried tephra in regards to EM1-like

source contributions suggests that Sr isotopes may be decoupled from Pb and Nd isotopes

and that the EM1-like source is in fact relatively minor at Thríhnúkagígur.

The lack of a systematic trend in any of the RP fissure swarms towards the EM1-

like component on the 87Sr/86Sr vs. 143Nd/144Nd (Figure 3.13) or 206Pb/204Pb vs. 87Sr/86Sr

(Figure 3.14) plots lends support to the hypothesis that altered crust contributes to the

genesis of Reykjanes Peninsula magmas (Gee et al., 1998a). A crustal component

hydrothermally altered by seawater, as given by Gee et al. (1998a) plots near the EM1-

like component (Figure 3.13). The broad array defined by Icelandic lavas in the Sr-Nd

isotopes compared to the tight array in Pb-Pb isotopes is consistent with variable

contributions to mantle-derived magmas by a highly radiogenic 87Sr/86Sr hydrothermally

seawater-altered crustal component. With 87Sr/86Sr ratios in excess of 0.7042 in drill

80

cores from modern RP geothermal systems (Elderfield and Greaves, 1981), even small

amounts of contamination to these lavas could produce notable changes to the parental Sr

isotopic composition. Hydrothermally altered crust (HAC) is likely to have a much

stronger effect on Sr isotopes than Nd or Pb isotopes because it is both fluid mobile and

there is a large and highly radiogenic reservoir of Sr in the ocean. So although a single

point is plotted for HAC on Figure 3.23, small contributions from seawater-altered crust

may shift magma compositions towards more radiogenic 87Sr/86Sr values at invariant

143Nd/144Nd ratios.

3.5.1.4 Summary

The anomalous Nb/U and Ce/Pb ratios in lavas on the Reykjanes Peninsula may

not be adequately explained by the three major sources proposed for Iceland. The

isotopically enriched plume source is common to all of Iceland and contributes more to

lavas in the southern EVZ, such as Hekla and Katla, than on the RP. Therefore, Nb/U or

Ce/Pb ratios would be expected to be higher in the southern EVZ than on the RP if it

were a signature of the enriched plume component. Based on the Pb isotope vs. Nb/U

plots (Figure 3.12), high Nb does not seem to be a characteristic of the depleted

component, which is partially defined by its depletion in trace elements, particularly Nb

(Thirlwall et al., 2004). An additional source with Pb isotopic compositions similar to the

enriched mantle plume source must be present beneath the RP, characterized by

unusually high Nb/U and Ce/Pb that sets it apart from the enriched component beneath

the Eastern Volcanic Zone.

81

Of the Sr-Nd-Pb isotope systematics, only Pb hints at the EM1-like component’s

contribution to Thríhnúkagígur lavas and dikes with slightly higher 208Pb/204Pb in some

samples for their 206Pb/204Pb ratios. This component does not contribute to the buried

tephras on the 206Pb/204Pb vs. 208Pb/204Pb plot, contradicting the Sr-Nd isotope plot where

the high 87Sr/86Sr component contributed equally to the buried tephra as to the lavas and

dikes. It is therefore unlikely that the EM1-like source is responsible for the highly

radiogenic Sr isotope ratios. Similarly, the sample (TNG-14-24) with the seemingly

largest EM1-like signature on the Pb-Pb plot, has the least radiogenic Sr composition,

further demonstrating that the Sr isotopes are decoupled from Pb and Nd isotopes. The

highly radiogenic 87Sr/86Sr ratios observed at Thríhnúkagígur suggest a basaltic crustal

component that has been hydrothermally altered by seawater.

On the basis of radiogenic isotopes and Nb/U ratios, there may be up to 5 mantle

sources that contribute to Thríhnúkagígur magmas as well as a hydrothermally altered

crustal component. These are 1) the depleted source, 2) the EM1-like source, 3) the

enriched plume source with Nb/U ratios more typical of MORB and OIB lavas, 4) the

enriched plume source with enriched Nb, and 5) the source proposed by Thirlwall et al.

(2004) similar to the depleted source, but with 87Sr/86Sr ratios between the depleted and

the EM1-like source. It is more likely that there are 2 mantle source domains – the

depleted source common to all of southern Iceland and Nb-enriched plume source unique

to the Reykjanes Peninsula – that are significant in generating Thríhnúkagígur magmas,

possibly with minor contributions from an EM1-like source. A fourth component

characterized by high 87Sr/86Sr that is decoupled from Nd and Pb isotope values likely has

its origins in the crust.

82

3.4.2 Magmatic processes

3.5.2.1 Fractional crystallization

The samples from Thríhnúkagígur form a nearly horizontal trend on a plot of

CaO/Al2O3 vs. MgO (Figure 3.4), which indicates olivine-dominated fractional

crystallization and removal in this system. The slight decrease in CaO/Al2O3 with

decreasing MgO in the buried tephra is consistent with removal of clinopyroxene. Some

lava and dike samples show a range to higher Ca/Al2O3 over a limited range of MgO,

which suggests that plagioclase may have been removed from these samples.

Fractional crystallization is evident from the major and trace element variations

with MgO in Thríhnúkagígur lavas and dikes as well as in the buried tephra. Both Ni and

Cr show a positive correlation with MgO, indicating that olivine is the dominant mafic

phase during fractional crystallization (Figure 3.4), as these elements are all compatible

in olivine. Scandium, which is compatible in clinopyroxene, decreases with decreasing

MgO in the buried tephras (Figure 3.4). Conversely, there is a negative trend in Sc with

MgO in the Thríhnúkagígur lavas and dikes, which is consistent with an increase of

incompatible elements in the residual magma as olivine (but not clinopyroxene) is

partitioned out of the system.

Strontium and europium are generally incompatible in mafic phases, but are

compatible in anorthitic plagioclase. Strontium anomalies behave similarly to those in Eu

(Figure 3.5), but whereas the Eu anomalies were subtle if present, the negative Sr

anomaly in the buried tephra and the positive anomalies in the lavas and dikes are

obvious on multielement diagrams (Figure 3.5). Primitive mantle normalized Sr values

83

converge on a relatively narrow range of values (8.37-12.1) across sample suites

compared to other elements. Plagioclase accumulation or removal could possibly explain

the variations in the Sr and Eu data. Alternatively, these data could be interpreted as

being consistent with lithospheric buffering of magma before its injection into the crust.

If the melt was generated in or approached equilibrium with plagioclase-bearing

peridotite in the lithospheric mantle prior to intruding the crust (Gee et al., 1998a), a more

discrete range of Sr and Eu would be expected. Regardless of possible lithospheric

interaction, the Eu/Eu* values close to unity suggest that plagioclase fractionation plays a

relatively small role in the evolution of lavas, dike, and buried tephras.

3.5.2.2 Assimilation of hydrothermally altered basaltic crust

Hydrothermal alteration by seawater shifts basaltic compositions towards more

radiogenic 87Sr/86Sr ratios. Previous studies have used Sr isotope ratios to argue that

assimilation of crust altered by seawater contributes in excess of 20% to magmas in south

Iceland during their residence in the crust (e.g. Sigmarsson et al., 1991; Sigmarsson et al.,

1992; Gee et al., 1998a). Oxygen has also been used to investigate the assimilation of

altered crust (Bindeman et al., 2006; Bindeman et al., 2008), because the major element

oxygen is a more robust tracer than trace-level Sr. But even oxygen isotopes are not a

well-agreed upon proxy. Other authors alternatively interpret anomalously light oxygen

isotope ratios to be a reflection of a recycled oceanic crustal component in the mantle

plume beneath Iceland (Kokfelt et al., 2006).

Hydrothermally altered crust contains secondary and hydrated minerals that give

it a lower density than unaltered basalt, and perhaps a lower density than basaltic

84

magmas. This density contrast can cause the ascending magma to stall and form sill-like

bodies (Taisne et al., 2011). Mechanisms that may allow for continued ascent include

increasing the pressure in the magma body, magmatic etching of a pathway through the

altered crust, or the assimilation of altered crust until the density of the magma is lowered

enough to continue to ascend buoyantly. Given that the altered crust plays a role in

stalling magmas in the crust, perhaps it should be expected that there is some amount of

interaction with a crustal component that may contribute to Icelandic magmas.

Nevertheless, there are few geochemical tools available to unambiguously trace

the bulk assimilation of basaltic crust in basaltic magma. The preponderance of seawater-

dominated hydrothermal systems on the Reykjanes Peninsula allows for 87Sr/86Sr ratios to

be used to model contamination of magmas in this part of Iceland (Raffone et al., 2008;

Marks et al., 2015). This is because seawater has much more radiogenic Sr than Icelandic

basalts and it has been shown to impart ratios as high as 0.7042 on altered basalts in the

Reykjanes fissure system (Elderfield and Greaves, 1981). If the Thríhnúkagígur Sr

isotopes are assumed to be derived partially from a hydrothermally altered basaltic

component at depth, a maximum estimate of crustal contamination can be determined.

This model builds off of the work of previous studies that use a hydrothermally altered

crustal component to explain the range of 87Sr/86Sr and 18O in fresh basalts on the

Reykjanes Peninsula; they measure compositions extending to radiogenic values of

87Sr/86Sr = 0.703295, and very light values of δ18O = +3‰ on the Reykjanes Peninsula

(Gee et al., 1998a; Gee et al., 1998b; Peate et al., 2009).

In order to assess the possible contributions of hydrothermally altered crust,

potential mixing between Sr and Nd end members in the Thríhnúkagígur system are

85

modeled. The model in Figure 3.15 shows two binary mixing trends on a Sr-Nd isotope

plot. The first is a single, nearly horizontal mixing curve between dike sample TNG-14-

24 and a conservative hydrothermally altered crustal component proposed by Gee et al.

(1998a). The dike sample is chosen as a mixing end member because it has the least

radiogenic Sr isotopic composition (as well as the least radiogenic Pb and most

radiogenic Nd isotopes) of any of the lavas and dikes. The large range of 87Sr/86Sr ratios

reported in modern geothermal systems (Elderfield and Greaves, 1981) makes a ratio of

0.7036 a probable minimum estimate. The combination of the most depleted dike sample

and a minimum estimate for the crustal component will yield a maximum estimate for the

amount of assimilated hydrothermally altered crust in the suite of Thríhnúkagígur lavas

and dikes.

The mixing curve between the theoretical altered crustal composition and the dike

show that a maximum of 25% hydrothermally altered crust may explain the full range of

Thríhnúkagígur lavas and dike compositions in Sr-Nd isotope space. The majority of the

samples may be explained by less than 15% of the crustal component. It is important to

emphasize that these represent maximum estimates. True quantities of crustal

contamination could be much lower if the altered crust has a more radiogenic 87Sr/86Sr

value than 0.7036; which seems likely given the observed values in modern geothermal

systems (Elderfield and Greaves, 1981), or if there is a range of 87Sr/86Sr in the primary

mantle melts. Indeed, more extensive alteration by seawater could both drive the 87Sr/86Sr

to higher ratios and make it easier to assimilate by facilitating the growth of more

secondary and hydrous minerals. In this case small amounts of a hydrothermally altered

86

crustal component could explain the observed 87Sr/86Sr ratios at Thríhnúkagígur and on

the rest of the Reykjanes Peninsula.

3.5.2.3 Buried tephra as a component in Thríhnúkagígur lavas

Evidence from direct field relationships observed in the cave below

Thríhnúkagígur’s main cone suggests that the buried tephra was physically scoured out

and entrained by the magma that fed the eruption. One of the most convincing

observations is the presence of hollow, cylindrical proto-conduits that extend vertically

into the overlying tephra immediately adjacent to the cross-cutting dike. In order to

evaluate this direct observation of processes occurring within the upper magmatic

plumbing system, isotopic and trace element mixing models are used to evaluate the

tephra assimilation hypothesis geochemically.

The Sr-Nd isotope data in this study do not form a simple binary mixing trend

between the hydrothermally altered crust and the dike. In order to assess whether the

buried tephra could provide an additional crustal component distinct from the

hydrothermally altered crust, the buried tephra with the least radiogenic 87Sr/86Sr (TNG-

14-26) was chosen as a third component. Given the positioning of the buried tephra ~100

m below the surface, a two-step mixing process is assumed. First, variable mixing with

hydrothermally altered crust occurs at depths of a few kilometers. Then these variably

affected magmas pick up the buried tephra immediately prior to or during the eruption.

Thus, compositions reflecting variable mixtures between mantle-derived magma and

hydrothermally altered crust (the near-horizontal curve in Figure 3.15 described in the

previous section) are then mixed with the representative tephra component. All but a

87

single outlier require less than 30% of the buried tephra, which is consistent with the

maximum estimate of 29% obtained based on physical observations made in Chapter 2.

Immobile elements such as HFSEs and REEs can also be used to assess mixing

between mantle-derived magma and the buried tephra. Ratios of HSFEs (Nb/Zr) and

REEs (La/Yb) were chosen for two reasons. First, there is an observable contrast in these

ratios between the buried tephra and the Holocene lavas and dikes. Second, the use of

immobile trace elements helps to avoid the potential effects of low-temperature alteration

to the buried tephra since it became exposed at depth in the damp cave. Secondary

alteration could potentially overprint the systematics of fluid-mobile elements.

Three end members were chosen to represent possible original magma

compositions. The plugged conduit, TNG-14-18, is used because it has the highest Mg#

(61) of the sample suite. The chilled margin of the dike, TNG-14-27i, is used because of

its low immobile trace element ratios. Finally, the most isotopically depleted dike sample

used in the Sr-Nd modeling is implemented here for consistency. The average of five of

six buried tephra samples is used to represent the buried tephra end member. Sample

TNG-14-28 is excluded from this average because of its high CIA (58) and unique

sedimentary structures, which are found only in two 20 cm thick outcrops out of tens of

meters of exposed tephra in the cave. Therefore, it is not considered to be representative

of the average buried tephra. End member compositions are given in Table 3.6.

Refractory trace element ratios Nb/Zr and La/Yb (Figure 3.16) were used to

determine relative mixing proportions. Depending on the magmatic end member,

estimates of tephra addition vary considerably. The plugged conduit yields the highest

estimates of 28-35% tephra to explain lavas and dike compositions. The chilled margin

88

gives more moderate estimates of 12-18%; and the dike used as the end member in the

Sr-Nd modeling suggests 0-8% tephra addition to the erupting magma. The plugged

conduit (TNG-14-18) represents a slightly more primitive magma (MgO > 9%) than that

erupted during the majority of the eruption (MgO = 7.5-8%). It is not entirely consistent

with the major element systematics of the rest of the system (see for example MgO vs.

Al2O3 in Figure 3.3), and may result in an over-estimate of the inherited tephra

component. Thus the maximum estimate from immobile trace element modeling is

consistent with the maximum estimate of 29% buried tephra produced with volume

estimates of the cave from a LiDAR-derived DEM and the erupted products.

The variability of these three estimates, none of which overlap, does not lend

strong support to the hypothesis that the cave was formed by the erosion and entrainment

of buried tephra into the erupting magma. The dike (TNG-14-24) used in the Sr-Nd

isotope mixing model produces low estimate of tephra addition (0-8%). As this seems the

most reasonable primitive mixing end member based on its isotopes ratios (as well as its

wt% Al2O3 and Sc concentration; see Figures 3.2 and 3.3), the low estimate warrants

some consideration. It is possible given the unconsolidated nature of the buried tephra

that it may have been rapidly scoured, entrained, and erupted. The proximity of the

buried tephra cone to the surface would not necessarily allow much time for the magma

to have mixed entrained tephra homogeneously. Early lavas buried within the main cone

may contain the bulk of the buried tephra, while tephra-poor lavas and dikes formed later

in the eruption are the only ones currently available to sample.

The extent to which the buried tephra was entrained in the erupting magma is not

clear from geochemical mixing models. The Sr-Nd isotope mixing model benefits from

89

the addition of a tephra component. On the other hand, the immobile element mixing

model does not necessarily require a tephra component to explain the observed

compositions of the lavas and dikes. These lines of evidence, at best, weakly support the

hypothesized entrainment of tephra to form the cave. At worst, they are not inconsistent

with the hypothesis.

3.5.3 Brennisteinsfjöll plumbing system

Variations in lava geochemistry over the course of an eruption have been

documented elsewhere in Iceland in basaltic (Furman et al., 1991) and intermediate

systems (Sigmarsson et al., 2011). It is not unreasonable to suspect that a dike and sill

system propagating towards the surface might contain semi-isolated sill-like bodies at

different levels in the crust. These various pockets of magma may experience different

geochemical evolutionary histories. Further, different injections of melt to the shallow

crustal part of the system may be derived from different proportions of enriched and

deplete mantle components. The lack of a central homogenizing magma chamber at

Brennisteinsfjöll may facilitate production and/or preservation of a heterogeneous

magma, and by extension the smaller fissure swarm of Thríhnúkagígur.

The heterogeneous nature of the major and trace element and radiogenic isotope

geochemistry in the Thríhnúkagígur lavas and dikes reflects the lack of a large

homogenizing magma chamber in the magmatic plumbing system. Instead, a network of

interconnected sills and dikes could both preserve initial heterogeneity in mantle-derived

magmas and facilitate the development of additional heterogeneity by independent

evolution of small pockets of magma at multiple levels within the crust (Figure 3.17).

90

The absence of a central volcano in Brennisteinsfjöll is consistent with a conceptual

model of a network of small sill-like magma bodies as well as the intrusion and eruption

of magmas with variable geochemistry.

The top of the plumbing system exposed in the cave preserves moderate

variations in magma composition at different intervals in a single eruption. For example,

the plugged conduit is more than 1.5% higher in MgO than all other samples. The chilled

margin of the main dike sampled from the wall of the cave anomalously high silica

relative to the other unaltered Holocene samples (Table 3.2). These two samples also

have markedly different incompatible trace element ratios relative to other Holocene

lavas (e.g., lower Nb/Zr and La/Yb; Figure 3.16). It may be expected that these samples

would have the most depleted isotopic compositions as well, but this is not the case.

Their Sr-Nd-Pb isotopes are indistinguishable from the rest of the lava and dike samples.

Hydrothermally altered crust or hyaloclastites can be much less dense than

crystalline basaltic lavas and could stall ascending magmas at many intervals in the crust

(Taisne et al., 2011). The variable crustal assimilation estimates (0-25%) suggested by

the Sr isotopic ratios in the Thríhnúkagígur lavas are consistent with the co-eruption of

relatively unmixed magmas produced in separate sill-like intrusions. Different sills may

incorporate variable amounts of hydrothermally altered crust. Further, the altered material

itself likely has variable Sr isotopic compositions.

The independent evolution of many different pockets of magmas could allow for a

wide range of compositions in a single eruption. This model, in the absence of a central

homogenizing magma chamber at Brennisteinsfjöll, invokes an interconnected network

of sills and dikes that may facilitate the development of chemical heterogeneity in

91

different locations in the plumbing system. Once an eruption is triggered, these relatively

isolated pockets of magma are unable to become well-mixed and compositional

differences are preserve in the lavas and the shallow magmatic plumbing system.

This model is also consistent with observations made during and prior to the

eruption at Eyjafjallajökull. Although this stratovolcano is markedly different in nature

from the fissure swarms of the Reykjanes, the inflation or injection of a mafic sill-like

from a deep source is likely to be the dominant intrusive style on the RP, which lacks

large homogenizing magma chambers. The deformation at Eyjafjallajökull in 1994 and

1999 is modeled to have occurred at 3.5-6.5 km depth based on InSAR (Pedersen and

Sigmundsson, 2004; Pedersen and Sigmundsson, 2006; Sigmundsson et al., 2010), GPS

and tilt measurements (Sturkell and Sigmundsson, 2003), and seismic observations

(Tarasewicz et al., 2014). A small volume flank eruption presumably sourced from this

intrusion had a mafic composition. Interaction with a shallow more silicic chamber

sparked the summit eruption (Sigmundsson et al., 2010; Sigmarsson et al., 2011). As the

summit eruption progressed, the chemistry of the tephra changed composition hinting at

mixing of the mafic component with various more evolved components at different times

during ascent (Keiding and Sigmarsson, 2012). This observation is consistent with the

observed microseismic earthquake swarms concentrated at particular depths propagating

down from the middle crust to upper mantle with each renewal of eruptive explosivity

(Tarasewicz, Brandsdóttir, et al., 2012). Collectively, these observations have led

multiple authors to suggest a complex multi-sill magmatic plumbing system

(Sigmundsson et al., 2010; Keiding and Sigmarsson, 2012; Tarasewicz, White, et al.,

2012).

92

3.6 Conclusions

The lavas and dikes at Thríhnúkagígur have Sr-Nd-Pb isotopic ratios that fall

within the range of data previously reported for the Reykjanes Peninsula and at

Brennisteinsfjöll. These compositions can be explained by three component mixing of an

isotopically and Nb enriched plume component and a depleted component, with minor

contributions from an EM1-like component. The southern EVZ requires an isotopically

indistinguishable enriched plume component from the one contributing to the RP, but

without the unusually enriched Nb. Lavas and dikes at Thríhnúkagígur plot with more

enriched RP compositions with a minor EM1-like signature. Unlike the Holocene lavas

and dikes, the buried tephra does not have a discernable EM1-like contribution and

cluster at the extreme enriched end of previously defined RP compositions. Although the

buried tephra is more enriched than the Holocene lavas and dikes, the difference is small

compared to the full range of compositions expressed on the RP. Because the buried

tephra is ≥100 ka in age, this contrast may also reflect a gradual depletion or dilution of

the enriched component under Brennisteinsfjöll through time.

The 87Sr/86Sr ratios, like the Nd and Pb isotope ratios, are generally consistent

with Reykjanes Peninsula lavas, except that some samples have relatively high Sr isotope

ratios at low 206Pb/204Pb and 143Nd/144Nd ratios. An additional mantle source may be

invoked to explain this variation, but is only discernable on the 206Pb/204Pb vs. 87Sr/86Sr

plot. The necessity of this additional mantle end member is difficult to evaluate because

Sr isotopes seem to be decoupled from the Nd-Pb isotopes. The most EM1-like sample

on the 206Pb/204Pb vs. 208Pb/204Pb plot is the least EM1-like in Sr-Nd space. Therefore,

basaltic crust altered at depth in seawater-dominated hydrothermal systems is proposed as

93

an alternative source for radiogenic Sr. Seawater has highly radiogenic Sr, so even minor

contributions from the altered crust could affect 87Sr/86Sr ratios in erupted lavas.

The isotope, major, and trace element geochemistry of Thríhnúkagígur is similar

to other erupted compositions on the Reykjanes Peninsula. However, the unique cave

below the main vent of the 3500 BP eruption allows compositions from the shallow

plumbing system to be constrained in addition to the surface lavas. The intrusive samples

collected from dike exposures within the walls of the cave have a wider range in major

and trace element abundances than the lavas. This demonstrates magma heterogeneity

within the small volume of magma along its final ascent to the surface vent, which is not

apparent in the majority of surficial flows. It is possible that during the peak of

monogenetic fissure eruptions, the magma is well-mixed relative to the initial and final

stages, when flow rates are lower. We propose that even in relatively homogeneous

monogenetic eruptions, significant chemical variation may persist into the uppermost

plumbing system, essentially to the point of the eruption. The surface lavas that represent

this heterogeneity may be buried under subsequent lava flows or preserved only in very

localized deposits near the vent, or they may be mixed out during the eruptive event

itself.

The proposed magmatic plumbing system of interconnected dikes and sills stalled

at hydrothermally altered layers in the middle to upper crust below Brennisteinsfjöll

provides a mechanism that may both preserve variations in mantle derived magma

compositions and facilitate additional magma heterogeneity. This type of plumbing

system is consistent with geophysical observations at other volcanic systems in Iceland,

as are the crustal residence times. The time scales of mafic magma storage, in particular,

94

have implications for volcanic hazards. If deformation is identified, eruption could occur

tens of years later.

Observations made from within the cave at Thríhnúkagígur suggest that buried

tephra has been assimilated or simply excavated by the magma en route to the surface.

Immobile trace element and radiogenic isotopes mixing models are used to test the

conclusions of magmatic entrainment of the buried tephra presented in Chapter 2. These

models produce similar estimates for the maximum amount of tephra that could have

been entrained and erupted in the Thríhnúkagígur lavas. These models produce maximum

estimates of ~28% and ~35% respectively, and 0-8% at minimum. Variation in initial

magma compositions may partially explain the variation in the estimates from these

models. Another possible source of variability is the extent of buried tephra incorporation

as the eruption progresses. It is possible that during the initial phases of the eruption,

likely more explosive with fire fountaining along a fissure, the buried tephra may have

been blown out by the erupting magma. These lavas with a greater buried tephra

component are likely at the base of the main cone at Thríhnúkagígur. Accordingly, the

surficial lavas in this study may record less of a buried tephra component if the mass

balance approach based on geophysical evidence is considered to be robust.

The lack of consensus about how to identify uniquely crustal tracers of

contamination in Icelandic magmas remains a problem. Hydrated, palagonitized, or

vitreous basaltic material, akin to the buried tephra beneath Thríhnúkagígur, is likely to

have a lower melting temperature than holocrystalline basalt. Fine-grained, low-density,

poorly consolidated materials (e.g. hyaloclastites, tephras, tillites) are easily

disaggregated, increasing the probability of magma stalling, and have high surface area,

95

which may encourage rapid digestion or entrainment. Nonetheless, even at

Thríhnúkagígur, where there are reasonable physical constraints on the maximum mass of

basaltic crustal component assimilated and completely characterized geochemistry,

including the upper magmatic plumbing system in addition to surficial lava flow, mixing

models are unable to provide internally consistent estimates of assimilation. The ability

of magma compositions to preserve an unambiguous record of interaction with altered

basaltic crustal materials may be limited to systems involving extreme hydrothermal

alteration of the crust as recorded by stable isotopes and radiogenic seawater-derived

strontium.

96

Table 3.1. Mineral and groundmass modes in percent. Modes are normalized to 100% after removing the vesicle mode, which is reported

separately.

Sample ID Plagioclase Olivine Oxides* Groundmass/

Glass

Tephra/

Contaminant Clinopyroxene Vesicles

Dike

TNG-14-14m 10.9 4.4 13.5 70.4 0.8 19.2

TNG-14-24 18.5 2.9 78.6 13.5

TNG-14-25 40.5 7.0 10.1 0.0 42.4 21

TNG-14-27i 44.3 10.2 10.2 0.0 35.2 12

Lavas

TNG-14-32 18.3 4.9 59.8 17.1 18

TNG-14-33 17.8 5.6 10.5 61.4 4.0 0.6 24.5

TNG-14-34 15.4 6.3 74.8 3.6 36.5

TNG-14-37 15.8 6.1 70.1 7.2 0.7 27.5

Miðhnúkar

TNG-14-35 6.3 7.4 78.1 8.2 27.5

Vent Spatter

MC-612-03 11.7 1.3 87.0

Plugged Conduit

TNG-14-18 2.1 8.9 89.1 4

Tephra

TNG-14-15 5.8 2.9 91.3

TNG-14-17 10.0 2.9 87.1

TNG-14-19

TNG-14-22 10.7 1.3 88.0

TNG-14-26 10.2 2.6 87.2

TNG-14-28 3.6 96.4

97

Table 3.2. Major element chemical analyses in weight percent oxides. Data were collected by a Perkin-Elmer Optima 5300DV ICP-AES for a

representative sample set for the units present at Thríhnúkagígur. Mg# and the chemical index of alteration (CIA) are unitless numbers. Mg# is

defined as 100 * Mg/(Mg + Fe) where Mg and Fe are molar abundances. CIA is defined as 100 * Al2O3/(Al2O3 + CaO + Na2O + K2O). For plotting

major element compositions, all samples were normalized to 100% after removing LOI. *Duplicate analysis. Duplicates are averaged for all figures.

Sample ID TNG-14-

14m

TNG-14-

24

TNG-14-

25

TNG-14-

27e

TNG-14-

27e*

TNG-14-

27i

TNG-14-

18

TNG-14-

18*

TNG-14-

15

TNG-

14-17

TNG-14-

19

Sample Type Dike Dike Dike Dike Dike Dike Plugged

conduit

Plugged

conduit

Buried

tephra

Buried

tephra

Buried

tephra

Description Interior Interior Interior Chilled

margin

Chilled

margin Interior Rockfall Rockfall

Uncon-

solidated Bomb

Uncon-

solidated

Location Cave,

NNE end

Cave,

SSW end

Cave,

SSW end

Cave,

SSW end

Cave,

SSW end

Cave,

SSW end

Cave, N

end

Cave, N

end

cave,

N/NNE

cave, N

end

cave,

NNW end

SiO2 48.11 47.42 47.11 48.20 49.25 47.23 47.58 47.68 48.38 48.11 48.18

TiO2 1.77 1.58 1.54 1.19 1.23 1.55 1.41 1.43 2.37 2.33 2.26

Al2O3 14.69 15.71 15.91 14.88 15.06 15.64 15.22 15.34 14.63 14.56 14.93

Fe2O3T 13.20 12.25 12.15 10.78 11.12 12.08 12.30 12.45 14.06 13.85 13.55

MnO 0.20 0.19 0.19 0.18 0.18 0.19 0.19 0.19 0.21 0.21 0.20

MgO 7.55 7.63 7.98 7.86 7.87 7.89 9.58 9.54 6.60 6.51 6.55

CaO 12.03 12.04 11.84 13.38 13.31 12.14 12.00 11.95 11.36 11.43 11.50

Na2O 2.12 2.13 2.15 2.04 2.09 2.10 1.95 2.00 2.30 2.32 2.28

K2O 0.18 0.13 0.14 0.10 0.08 0.15 0.10 0.07 0.28 0.28 0.30

P2O5 0.18 0.15 0.18 0.09 0.11 0.16 0.14 0.14 0.29 0.30 0.29

LOI 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.69

Total 100.03 99.26 99.22 98.69 100.32 99.12 100.47 100.81 100.53 99.94 100.74

Mg# 53 55 57 59 58 56 61 60 48 48 49

CIA 51 52 53 49 49 52 52 52 51 51 51

98

Table 3.2 continued.

Sample ID TNG-14-22 TNG-14-26 TNG-14-28 TNG-14-32 TNG-14-33 TNG-14-34 TNG-14-37 MC-612-03 TNG-14-35

Sample Type Buried

tephra

Buried

tephra

Buried

tephra Lava Lava Lava Lava

Vent

spatter

Miðhnúkar

lava

Description Uncon-

solidated

Uncon-

solidated

Palagon-

itized

Flank vent

flow

Main vent

flow

Main vent

flow

Main vent

flow

Main vent

scoria Lava flow

Location cave, NW

end

cave

SW/SSW

cave

SW/WSW

N 63°

59.865'

W 21°

42.015'

N 63°

59.958'

W 21°

41.880'

N 63°

59.972'

W 21°

41.658'

N 63°

59.836'

W 21°

42.098'

top of main

vent

N 63°

59.981'

W 21°

41.455'

SiO2 (%) 47.95 47.81 46.70 48.11 47.26 47.33 47.70 49.22 46.34

TiO2 (%) 2.29 2.35 1.89 1.65 1.60 1.60 1.58 1.71 1.82

Al2O3 (%) 14.77 14.42 16.09 15.14 15.66 15.69 15.60 16.15 14.04

Fe2O3T (%) 14.09 13.84 13.45 12.73 12.29 12.28 12.20 13.04 13.55

MnO (%) 0.20 0.21 0.17 0.19 0.19 0.19 0.19 0.18 0.21

MgO (%) 6.20 6.76 6.03 7.66 7.66 7.67 7.74 7.21 8.62

CaO (%) 10.84 11.33 9.79 11.94 12.01 12.00 11.89 9.91 10.93

Na2O (%) 1.95 2.29 1.65 2.07 2.05 2.09 2.07 1.65 2.01

K2O (%) 0.31 0.29 0.20 0.14 0.15 0.16 0.15 0.11 0.19

P2O5 (%) 0.29 0.28 0.24 0.16 0.15 0.16 0.17 0.17 0.17

LOI(900C) 1.11 0.00 2.79 0.00 0.00 0.00 0.00 0.00 0.00

Total 99.99 99.62 99.00 99.82 99.03 99.17 99.29 99.37 98.34

Mg# 47 49 47 54 56 55 56 52 56

CIA 53 51 58 52 52 52 53 58 52

99

Table 3.3. CIPW norms. Calculated from major element data and trace element concentrations of Sr, Ba, Cr, Ni, and Zr. Samples TNG-14-22,

TNG-14-28, and MC-612-03 are excluded from this table because they have been altered to an extent that makes their bulk chemistry fall outside

the range of compositions for which norms can be calculated.

Sample ID TNG-

14-14m

TNG-

14-24

TNG-

14-25

TNG-

14-27e

TNG-

14-27i

TNG-

14-18

TNG-

14-15

TNG-

14-17

TNG-

14-19

TNG-

14-26

TNG-

14-32

TNG-

14-33

TNG-

14-34

TNG-

14-37

Sample Type Dike Dike Dike Dike Dike Plugged

conduit

Buried

tephra

Buried

tephra

Buried

tephra

Buried

tephra

Flank

vent

lava

Main

vent

lava

Main

vent

lava

Main

vent

lava

Plagioclase 48.49 51.83 52.52 49.48 51.62 49.46 48.58 48.76 49.46 48.33 49.79 51.43 51.62 51.24

Orthoclase 1.06 0.83 0.83 0.53 0.89 0.53 1.65 1.65 1.83 1.77 0.83 0.95 0.95 0.95

Diopside 23.65 21.55 20.25 28.34 22.07 21.24 21.43 22.01 21.39 21.94 22.07 21.35 21.30 20.93

Hyper-sthene 15.07 12.06 10.53 11.50 10.35 11.17 18.61 17.50 17.19 16.82 16.39 12.89 12.37 14.71

Olivine 4.98 7.60 9.80 5.14 9.05 11.87 1.43 1.48 2.15 2.88 4.56 7.20 7.57 6.03

Ilmenite 3.40 3.06 2.89 2.34 3.00 2.70 4.54 4.48 4.33 4.54 3.17 3.10 3.10 3.06

Magnetite 2.90 2.71 2.70 2.42 2.68 2.70 3.07 3.06 2.97 3.06 2.80 2.73 2.73 2.7

Apatite 0.42 0.37 0.42 0.23 0.37 0.32 0.67 0.70 0.67 0.65 0.39 0.37 0.39 0.39

Zircon 0.01 0.01 0.01 0.01 0.01 0.01 0.03 0.03 0.03 0.03 0.01 0.01 0.01 0.01

Chromite 0.06 0.07 0.07 0.07 0.07 0.09 0.03 0.03 0.03 0.03 0.07 0.06 0.06 0.06

100

Table 3.4. Trace element chemical analyses. Data were collected by a Thermo X-Series II Quadruple ICP-

MS for samples from the Thríhnúkar fissure system. All concentrations are in ppm. NR = not reported.

*Duplicate analysis.

Sample

ID

TNG-

14-14m

TNG-

14-24

TNG-

14-25

TNG-

14-27e

TNG-

14-27e

TNG-

14-27i*

TNG-

14-18

TNG-

14-18*

Sample

Type Dike Dike Dike Dike Dike Dike

Plugged

conduit

Plugged

conduit

Li NR 4.24 4.41 NR 3.91 NR NR 4.93

Sc 44.2 42.5 40.8 53.4 55.8 41.7 39.9 42.7

V 382 352 342 380 391 347 314 338

Cr 295 314 331 318 317 333 382 391

Co 74.4 75.3 95.9 63.1 67.7 67.6 73.2 79.5

Ni 93.3 113 123 74.8 75.6 117 187 191

Cu 116 167 96.6 79.9 85.0 119 92.5 98.8

Zn 102 97.3 96.2 72.2 74.8 93.6 92.4 95.3

Rb 3.71 3.17 3.14 1.94 1.90 3.26 1.89 1.88

Sr 189 208 208 183 189 203 166 187

Y 25.7 23.7 22.1 17.5 17.2 23.0 22.2 22.3

Zr 87.6 78.8 78.5 50.3 50.0 77.6 62.7 62.1

Nb 11.7 10.6 10.8 6.15 6.27 10.2 6.64 6.56

Cs 0.09 0.10 0.10 0.07 0.10 0.07 0.05 0.08

Ba 60.9 52.5 51.7 39.1 36.6 55.9 33.0 33.5

Hf 2.31 2.05 2.01 1.46 1.41 2.06 1.72 1.66

Ta 0.83 0.66 0.65 0.45 0.38 0.73 0.48 0.41

Pb 0.54 0.49 0.39 0.30 0.30 0.46 0.26 0.26

Th 0.48 0.42 0.41 0.25 0.24 0.43 0.25 0.25

U 0.15 0.13 0.13 0.08 0.07 0.13 0.08 0.08

La 7.77 6.97 6.81 4.41 4.29 7.01 4.87 4.88

Ce 18.8 17.0 16.4 10.9 10.6 17.0 12.7 12.6

Pr 2.77 2.37 2.25 1.65 1.51 2.48 1.95 1.84

Nd 12.7 11.6 10.8 7.80 7.57 11.3 9.24 9.42

Sm 3.57 3.25 2.99 2.34 2.21 3.18 2.77 2.76

Eu 1.29 1.19 1.11 0.94 0.90 1.17 1.04 1.03

Gd 4.19 3.78 3.48 2.84 2.71 3.75 3.37 3.32

Tb 0.71 0.64 0.60 0.49 0.47 0.64 0.59 0.58

Dy 4.50 4.01 3.77 3.05 2.93 4.00 3.79 3.72

Ho 0.92 0.85 0.79 0.63 0.61 0.83 0.78 0.78

Er 2.61 2.40 2.21 1.80 1.72 2.35 2.24 2.23

Tm 0.39 0.35 0.32 0.26 0.24 0.35 0.33 0.33

Yb 2.44 2.24 2.09 1.66 1.59 2.17 2.07 2.08

Lu 0.38 0.33 0.31 0.26 0.24 0.34 0.33 0.31

101

Table 3.4 continued.

Sample ID TNG-14-15 TNG-14-17 TNG-14-19 TNG-14-22 TNG-14-26 TNG-14-28

Sample Type Buried

tephra

Buried

tephra

Buried

tephra

Buried

tephra

Buried

tephra

Buried

tephra

Li 5.00 4.98 NR NR 5.19 NR

Sc 43.8 44.3 43.4 41.7 48.5 40.4

V 403 408 406 387 445 343

Cr 139 141 143 136 160 224

Co 66.6 107 59.4 51.9 69.2 62.2

Ni 75.9 73.5 77.9 67.4 80.8 136

Cu 317 187 180 265 190 271

Zn 116 117 114 113 128 98.4

Rb 6.33 6.34 6.67 8.16 6.99 4.07

Sr 248 256 241 236 280 185

Y 33.5 31.4 30.1 35.2 36.5 24.7

Zr 132 133 131 130 145 97.8

Nb 21.8 21.7 21.2 20.9 23.9 14.4

Cs 0.17 0.18 0.15 0.20 0.18 0.13

Ba 93.8 94.5 96.3 95.0 103 75.4

Hf 3.24 3.31 3.34 3.30 3.54 2.55

Ta 1.31 1.33 1.48 1.44 1.42 1.00

Pb 0.81 0.85 0.81 0.86 0.65 0.63

Th 0.84 0.86 0.85 0.85 0.91 0.65

U 0.25 0.26 0.27 0.27 0.26 0.18

La 13.5 12.9 12.9 13.2 14.5 9.11

Ce 33.4 31.2 31.0 32.0 34.4 21.9

Pr 4.58 4.24 4.41 4.55 4.56 3.13

Nd 22.0 19.9 19.6 20.0 21.3 14.1

Sm 5.56 5.05 5.00 5.18 5.47 3.78

Eu 1.86 1.73 1.73 1.78 1.87 1.32

Gd 6.18 5.46 5.47 5.82 5.90 4.22

Tb 1.00 0.90 0.90 0.96 0.97 0.71

Dy 5.97 5.39 5.41 5.85 5.79 4.40

Ho 1.24 1.10 1.08 1.19 1.18 0.90

Er 3.40 3.05 3.03 3.32 3.27 2.53

Tm 0.48 0.44 0.44 0.48 0.46 0.38

Yb 3.06 2.82 2.70 2.94 2.99 2.37

Lu 0.45 0.41 0.42 0.45 0.43 0.37

102

Table 3.4 continued.

Sample ID TNG-14-

32

TNG-14-

33

TNG-14-

34

TNG-14-

37

MC-612-

03

TNG-14-

35

Sample

Type Lava Lava Lava Lava

Vent

spatter Miðhnúkar

Li 5.31 NR NR NR 4.68 NR

Sc 46.7 41.7 42.0 40.1 45.9 41.6

V 387 354 354 342 385 382

Cr 326 297 294 287 315 319

Co 99.0 76.4 164 170 72.8 84.1

Ni 120 114 114 110 117 161

Cu 176 159 149 150 159 149

Zn 105 97.7 96.5 92.5 107 109

Rb 3.52 3.48 3.54 3.29 3.64 4.19

Sr 221 192 193 194 179 176

Y 25.3 24.3 24.0 23.2 23.4 27.4

Zr 87.3 82.4 82.5 77.7 92.7 97.9

Nb 11.9 11.0 11.0 10.4 12.5 12.5

Cs 0.11 0.08 0.08 0.08 0.13 0.11

Ba 57.3 56.2 56.0 56.2 55.6 61.0

Hf 2.23 2.18 2.18 2.07 2.33 2.57

Ta 0.71 0.78 0.78 0.74 0.76 0.88

Pb 0.55 0.61 0.91 0.50 2.40 1.38

Th 0.46 0.46 0.48 0.43 0.52 0.55

U 0.14 0.15 0.15 0.13 0.16 0.18

La 7.65 7.54 7.24 7.08 7.65 8.28

Ce 18.6 18.4 17.9 17.3 18.5 20.5

Pr 2.58 2.68 2.63 2.53 2.53 3.00

Nd 12.5 12.3 12.0 11.5 12.2 13.8

Sm 3.42 3.43 3.37 3.23 3.27 3.87

Eu 1.26 1.24 1.25 1.20 1.19 1.37

Gd 4.01 4.00 3.93 3.83 3.81 4.46

Tb 0.68 0.68 0.68 0.65 0.65 0.77

Dy 4.29 4.31 4.20 4.08 3.99 4.81

Ho 0.90 0.88 0.87 0.85 0.84 0.99

Er 2.53 2.50 2.46 2.38 2.34 2.80

Tm 0.37 0.37 0.37 0.35 0.34 0.42

Yb 2.36 2.34 2.33 2.21 2.17 2.64

Lu 0.35 0.36 0.36 0.35 0.33 0.41

103

Table 3.5. Pb-Sr-Nd isotope ratios. Data were collected by Finnigan MAT 262 multi-collector thermal ionization mass spectrometer (TIMS). Two

standard deviations are given in ten-thousandths for Pb ratios and one-millionths for 87Sr/86Sr and 143Nd/144Nd.

Sample ID TNG-14-

14m TNG-14-24 TNG-14-25

TNG-14-

27e TNG-14-27i TNG-14-18 TNG-14-15 TNG-14-17 TNG-14-19

Sample Type Dike Dike Dike Dike Dike Plugged

conduit

Buried

tephra

Buried

tephra

Buried

tephra

206Pb/204Pb 18.9042 18.5182 18.8017 18.7945 18.8128 18.6805 18.9241 18.6840 18.9437

2σ 228 112 224 55 36 82 77 28 69

207Pb/204Pb 15.5892 15.5525 15.6302 15.4768 15.5031 15.5237 15.5355 15.5346 15.5381

2σ 192 97 184 44 30 72 63 23 57

208Pb/204Pb 38.5980 38.2656 38.5401 38.3135 38.3958 38.3395 38.5259 38.3471 38.5224

2σ 463 235 456 108 74 167 154 57 139

87Sr/86Sr 0.703187 0.703135 0.703143 0.703203 0.703141 0.703088 0.703174 0.703186 0.703223

2σ 7 7 8 9 8 30 7 7 15

143Nd/144Nd 0.513031 0.513040 0.513036 0.513021 0.513029 0.513039 0.513012 0.513010 0.513010

2σ 8 15 6 10 20 8 8 8 6

104

Table 3.5 continued.

Sample ID TNG-14-22 TNG-14-26 TNG-14-28 TNG-14-32 TNG-14-33 TNG-14-34 TNG-14-37 MC-612-03 TNG-14-35

Sample Type Buried

tephra

Buried

tephra

Buried

tephra Flank lava

Main vent

lava

Main vent

lava

Main vent

lava Vent Spatter

Miðhnúkar

lava

206Pb/204Pb 18.9328 18.8634 18.8472 18.8418 18.8233 18.5738 18.8520 18.2889 18.4774

2σ 46 46 64 69 78 14 62 21 69

207Pb/204Pb 15.5411 15.4893 15.5191 15.5375 15.5338 15.5647 15.5352 15.5947 15.5628

2σ 37 38 52 56 62 11 54 18 58

208Pb/204Pb 38.5464 38.3865 38.4452 38.4790 38.4478 38.3561 38.4755 38.1947 38.2762

2σ 92 95 13 143 155 27 135 45 142

87Sr/86Sr 0.703305 0.703157 0.703191 0.703170 0.703213 0.703281 0.703178 0.703197 0.703184

2σ 8 8 10 7 9 8 8 6 7

143Nd/144Nd 0.513018 0.512991 0.513050 0.513000 0.513030 0.513044 0.513024 0.513022 0.513022

2σ 10 9 16 13 9 8 7 10 8

105

Table 3.6. Mixing end member compositions for Figures 3.14 and 3.15. HAC = hydrothermally altered crust; from Gee et al. (1998a).

End Member 87Sr/86Sr 143Nd/144Nd Sr (ppm) Nd (ppm) La/Yb Sm/Nd Nb/Zr

HAC 0.7036 0.51305 300 10.6

TNG-14-24 0.703135 0.513040 208 11.6 3.11 0.281 0.134

TNG-14-26 0.703157 0.512991 280 9.33 4.84 0.257 0.164

TNG-14-18 0.703188 0.513039 177 21.3 2.35 0.297 0.106

Avg. buried

tephra 4.26 0.256 0.163

106

Figure 3.1. Photomicrographs of two phenocryst population, a more tabular euhedral one (a) and a high

aspect ratio one (b) in lava TNG-14-14m; the phenocryst-poor chilled margin (c; TNG-14-27e) and flow

banding with interstitial clinopyroxene in the interior the same dike (d; TNG-14-27i); aphanitic,

phenocryst-poor plugged conduit, TNG-14-18 (e); and large olivine phenocrysts from the Miðhnúkar lava

flow, TNG-14-35 (f).

107

Figure 3.2. Total alkali-silica plot (TAS). All Thríhnúkagígur samples analyzed are subalkaline basalts.

Buried tephra samples are represented by triangle; lava and dike samples from the Holocene eruption are

circles; the lava sample from Miðhnúkar is a cross. TAS diagram from Le Bas and Le Maitre (1986).

108

Figure 3.3. Fenner diagrams of MgO vs. the major oxides and chemical index of alteration (CIA). Circles

represent lavas and dikes. Triangles represent buried tephra samples. The cross is the Miðhnúkar lava.

Open symbols indicate samples with a high CIA samples likely to be affected by secondary alteration.

109

Figure 3.4. Plots of MgO vs. trace elements and CaO/Al2O3. Circles represent lavas and dikes while

triangles represent buried tephra samples. The cross is the lava from Miðhnúkar. Open symbols indicate

samples with a high CIA and likely to be affected by secondary alteration.

110

Figure 3.5. Primitive mantle normalized multielement diagram for the main sample suites and unique

samples from Thríhnúkagígur. All samples are characterized by small positive Ba and U, and relatively

large Nb and Ta anomalies. All samples have moderate to large negative Pb anomalies, except for the

vent spatter (MC-612-03), which is thought to be a result of contamination. Most other trace elements

display similar patterns across regardless of sample type, except for Sr which has a positive anomaly in

relatively more depleted plugged conduit (TNG-14-18) and chilled margin of the dike (TNG-14-27e) and

a negative anomaly in the more enriched buried tephra. Primitive mantle normalization values are from

Sun and McDonough (1989).

111

Figure 3.6. Chondrite normalized rare earth element (REE) spider diagram by rock type. All samples have

a gently sloping negative trend with higher LREE values than HREE values. The chilled margin of the

dike (TNG-14-27e) has a weak positive Eu anomaly. Chondrite normalization values are from Sun and

McDonough (1989).

112

Figure 3.7. Plots of 206Pb/204Pb vs. 208Pb/204Pb (a) and 207Pb/204Pb (b). Circles represent lavas and dikes

while triangles represent buried tephra samples. The cross is the lava from Miðhnúkar. Open symbols

indicate samples with a high CIA. Error bars represent either the in run instrumental error or the standard

reproducibility, whichever source of error is larger.

113

Figure 3.8. Plot of 87Sr/86Sr vs. 143Nd/144Nd. The buried tephras have consistently lower 143Nd/144Nd ratios

compared to the lavas and dikes across the same range of 87Sr/86Sr ratios. Three dike samples have the

least radiogenic Sr isotope ratios at the same relative radiogenic Nd isotope ratio. Circles represent lavas

and dikes while triangles represent buried tephra samples. The cross is the lava from Miðhnúkar. Open

symbols indicate samples with a high CIA. Error bars represent either the in run instrumental error or the

standard reproducibility, whichever source of error is larger.

114

Figure 3.9. (La/Yb)N vs. Nb/Zr. Reykjanes Peninsula and the Western Volcanic Zone lavas in SW Iceland

extend to similarly high Nb/Zr ratios as lavas in the Eastern Volcanic Zone, but at much lower (La/Yb)N

ratios. Thríhnúkagígur lavas and dikes are consistent other lavas from SW Iceland while the buried tephra

record some of the most enriched values in the region. Circles represent lavas and dikes while triangles

represent buried tephra samples. The cross is the lava from Miðhnúkar. Open symbols indicate samples

with a high CIA. Data for Vestmannaeyjar are from Furman et al. (1991), Kokfelt et al. (2006), and Peate

et al. (2010); EVZ data are from Kokfelt et al. (2006), Peate et al. (2010), and Manning and Thirlwall

(2014); WVZ data are from Kokfelt et al. (2006) and Koornneef et al. (2012); and data from the

Reykjanes Peninsula are from Kokfelt et al. (2006), Peate et al. (2009), and Koornneef et al. (2012).

115

Figure 3.10. Plots of Nb vs. Nb/U (a) and Ce vs. Ce/Pb (b). Global ocean island basalts and mid-ocean

ridge basalts yield Nb/U ratios of 47 ± 10 at all concentrations of Nb and Ce/Pb ratios of 25 ± 5 at all Ce

concentrations (Hofmann et al., 1986). Dashed lines show these ranges. Lavas in SW Iceland on the

Reykjanes Peninsula and in the WVZ yield much higher Nb/U and Ce/Pb than observed elsewhere, not

just in Iceland, but globally. Circles represent lavas and dikes while triangles represent buried tephra

samples. The cross is the lava from Miðhnúkar. Open symbols indicate samples with a high CIA. Data for

Vestmannaeyjar are from Kokfelt et al. (2006) and Peate et al. (2010); EVZ data are from Sigmarsson et

al. (1992), Kokfelt et al. (2006), Peate et al. (2010), and Manning and Thirlwall (2014); WVZ and RP

data are from Kokfelt et al. (2006) and Koornneef et al. (2012).

116

Figure 3.11. Plots of 206Pb/204Pb vs. 208Pb/204Pb with proposed mantle end members from Thirlwall et al.

(2004) and Kitagawa et al. (2008). South Iceland lavas form an array primarily between an enriched and a

depleted end member on both plots. Thríhnúkagígur samples are broadly consistent with composition

observed on the Reykjanes Peninsula and Brennisteinsfjöll. They are pulled slightly off the NHRL (Hart,

1984) towards an EM1-like end member. Data for the southern Reykjanes Ridge (SRR) are from

Thirlwall et al., 2004; Reykjanes, Krísuvík, and Brennisteinsfjöll data are from Thirlwall et al., 2004,

Kokfelt et al., 2006, and Peate et al., 2009; Hekla and Katla data are from Park, 1990 and Furman et al.,

1995; Vestmannaeyjar data are from Furman et al., 1991, Chauvel and Hémond, 2000, Kokfelt et al.,

2006, and Peate et al., 2010.

117

Figure 3.12. Nb/U plotted against 206Pb/204Pb (a), 208Pb/204Pb (b). Solid black triangles represent

Thríhnúkagígur buried tephra. Solid black circles represent Thríhnúkagígur lavas and dikes. Grey circles

represent other Reykjanes Peninsula lavas from Kokfelt et al. (2006).

118

Figure 3.13. A plot of 87Sr/86Sr vs. 143Nd/144Nd. Thríhnúkagígur samples are more consistent with the

Reykjanes fissure swarm and Vestmannaeyjar than Brennisteinsfjöll and are influenced by an EM1-like

component, or more likely, HAC. Sr-Nd isotopes, data for the southern Reykjanes Ridge (SRR) are from

Thirlwall et al., 2004; Reykjanes, Krísuvík, and Brennisteinsfjöll data are from Thirlwall et al., 2004,

Kokfelt et al., 2006, and Peate et al., 2009; Hekla data are from Park, 1990, Sigmarsson et al., 1992, and

Furman et al., 1995; Katla data are from Park, 1990 and Furman et al., 1995; Vestmannaeyjar data are

from Furman et al., 1991 and Kokfelt et al., 2006.

119

Figure 3.14. A plot of 206Pb/204Pb vs. 87Sr/86Sr. Thríhnúkagígur samples are most consistent with the lavas

with highly radiogenic Sr isotopes from the Reykjanes fissure swarm. Data for the southern Reykjanes

Ridge (SRR) are from Thirlwall et al., 2004; Reykjanes, Krísuvík, and Brennisteinsfjöll data are from

Thirlwall et al., 2004, Kokfelt et al., 2006, and Peate et al., 2009; Hekla data are from Park, 1990,

Sigmarsson et al., 1992, and Furman et al., 1995; Katla data are from Park, 1990 and Furman et al., 1995;

Vestmannaeyjar data are from Furman et al., 1991 and Kokfelt et al., 2006.

120

Figure 3.15. Sr-Nd isotope mixing model for Thríhnúkagígur lavas and dikes. First, the dike sample with

the least radiogenic Sr isotope composition (TNG-14-24, purple circle) is mixed with a strongly

radiogenic hydrothermally altered crustal end member from Gee et al., 1998 (blue cross). Then mixing

occurs between various points along the first curve and the least radiogenic buried tephra sample, marked

with a red triangle.

121

Figure 3.16. Incompatible trace element ratio mixing curve, Nb/Zr vs. La/Yb. The average buried tephra

(red diamond) of the representative samples forms the enriched end member. The geochemistry of the

intrusive part of the system is heterogeneous, so three possible end members are used (yellow diamonds):

the plugged conduit (TNG-14-18), the chilled margin of the dike (TNG-14-27e), and isotopically most

depleted dike sample (TNG-14-24) used as an end member in Figure 3.13. For the latter two primitive

end members, green crosses on the mixing curves represent 20% buried tephra, blue crosses represent

10%, and red crosses represent 5%. Dikes and lavas representing possible mixtures of the buried tephra

and original magma composition are blue and yellow circles.

Figure 3.17. Cartoon of the magmatic plumbing system below Brennisteinsfjöll. The depth to the base of

the crust is to scale, but the volcanic features, dikes, and sills are not to scale. Hydrothermally altered

crust or hyaloclastite layers (HC) likely play a role in slowing magma ascent. Orange stippled zones

represent sills or sill-like intrusions that have stalled at low density layers.

References

Bindeman I., Gurenko A., Sigmarsson O. and Chaussidon M. (2008) Oxygen isotope

heterogeneity and disequilibria of olivine crystals in large volume Holocene basalts

from Iceland: Evidence for magmatic digestion and erosion of Pleistocene

hyaloclastites. Geochim. Cosmochim. Acta 72, 4397–4420.

Bindeman I., Sigmarsson O. and Eiler J. M. (2006) Time constraints on the origin of

large volume basalts derived from O-isotope and trace element mineral zoning and

U-series disequilibria in the Laki and Grímsvötn volcanic system. Earth Planet. Sci.

Lett. 245, 245–259.

Chauvel C. and Hémond C. (2000) Melting of a complete section of recycled oceanic

crust: Trace element and Pb isotopic evidence from Iceland. Geochemistry,

Geophys. Geosystems 1, n/a–n/a.

Clifton A. E., Sigmundsson F., Feigl K. L., Gunnar Guðmundsson and Árnadóttir T.

(2002) Surface effects of faulting and deformation resulting from magma

accumulation at the Hengill triple junction, SW Iceland, 1994-1998. J. Volcanol.

Geotherm. Res. 115, 233–255.

Condomines M., Muehlenbachs K., O’Nions R. K., Óskarsson N. and Oxburgh E. R.

(1983) Helium, oxygen, strontium and neodynium isotopic relationships in Icelandic

volcanics. Earth Planet. Sci. Lett. 66, 125–136.

Elderfield H. and Greaves M. J. (1981) Strontium isotope geochemistry of Icelandic

geothermal systems and implications for sea water chemistry. Geochim. Cosmochim.

Acta 45, 2201–2212.

Fronval T. and Jansen E. (1997) Eemian and Early Weichselian (140–60 ka)

Paleoceanography and paleoclimate in the Nordic Seas with comparisons to

Holocene conditions. Paleoceanography 12, 443.

Furman T., Frey F. a. and Park K.-H. (1991) Chemical constraints on the petrogenesis of

mildly alkaline lavas from Vestmannaeyjar, Iceland: the Eldfell (1973) and Surtsey

(1963?1967) eruptions. Contrib. to Mineral. Petrol. 109, 19–37.

Furman T., Frey F. A. and Park K.-H. (1995) The scale of source heterogeneity beneath

the Eastern neovolcanic zone, Iceland. J. Geol. Soc. London. 152, 997–1002.

Gee M. A. M., Taylor R. N., Thirlwall M. F. and Murton B. J. (2000) Axial magma

reservoirs located by variation in lava chemistry along Iceland’s mid-ocean ridge.

Geology 28, 699–702.

124

Gee M. A. M., Taylor R. N., Thirlwall M. F. and Murton B. J. (1998) Glacioisostacy

controls chemical and isotopic characteristics of tholeiites from the Reykjanes

Peninsula, SW Iceland. Earth Planet. Sci. Lett. 164, 1–5.

Gee M. A. M., Thirlwall M. F., Taylor R. N., Lowry D. and Murton B. J. (1998) Crustal

Processes: Major Controls on Reykjanes Peninsula Lava Chemistry, SW Iceland. J.

Petrol. 39, 819–839.

Hanan B. B. and Schilling J.-G. (1997) The dynamic evolution of the Iceland mantle

plume: the lead isotope perspective. Earth Planet. Sci. Lett. 151, 43–60.

Hart S. R. (1984) A large-scale isotope anomaly in the Southern Hemisphere mantle.

Nature 309, 753–757.

Higgins M. D. and Roberge J. (2007) Three magmatic components in the 1973 eruption

of Eldfell volcano, Iceland: Evidence from plagioclase crystal size distribution

(CSD) and geochemistry. J. Volcanol. Geotherm. Res. 161, 247–260.

Hofmann A. W., Jochum K. P., Seufert M. and White W. M. (1986) Nb and Pb in

oceanic basalts: new constraints on mantle evolution. Earth Planet. Sci. Lett. 79, 33–

45.

Jakobsson S. P., Jonsson J. and Shido F. (1978) Petrology of the Western Reykjanes,

Peninsula. J. Petrol. 19, 669–705.

Keiding J. K. and Sigmarsson O. (2012) Geothermobarometry of the 2010

Eyjafjallajökull eruption: New constraints on Icelandic magma plumbing systems. J.

Geophys. Res. 117, 1–15.

Kitagawa H., Kobayashi K., Makishima A. and Nakamura E. (2008) Multiple pulses of

the mantle plume: Evidence from tertiary icelandic lavas. J. Petrol. 49, 1365–1396.

Koerberle R., Hudak M. and Furman T. (2015) Crystal size distribution and textural

analysis at Thríhnúkagígur volcano, Iceland. In Geological Society of America

Annual Meeting p. 179.

Kokfelt T. F., Hoernle K., Hauff F., Fiebig J., Werner R. and Garbe-Schönberg D. (2006)

Combined Trace Element and Pb-Nd-Sr-O Isotope Evidence for Recycled Oceanic

Crust (Upper and Lower) in the Iceland Mantle Plume. J. Petrol. 47, 1705–1749.

Koornneef J. M., Stracke A., Bourdon B., Meier M. A., Jochum K. P., Stoll B. and

Grönvold K. (2012) Melting of a Two-component Source beneath Iceland. J. Petrol.

53, 127–157.

125

Le Bas M. J. and Le Maitre R. W. (1986) A chemical classification of volcanic rocks

based on the total alkali-silica diagram. J. Petrol. 27, 745–750.

LaFemina P. C., Dixon T. H., Malservisi R., Árnadóttir T., Sturkell E., Sigmundsson F.

and Einarsson P. (2005) Geodetic GPS measurements in south Iceland: Strain

accumulation and partitioning in a propagating ridge system. J. Geophys. Res. 110,

B11405.

Manning C. J. and Thirlwall M. F. (2014) Isotopic evidence for interaction between

Öræfajökull mantle and the Eastern Rift Zone, Iceland. Contrib. to Mineral. Petrol.

167, 959.

Marks N., Zierenberg R. A. and Schiffman P. (2015) Strontium and oxygen isotopic

profiles through 3 km of hydrothermally altered oceanic crust in the Reykjanes

Geothermal System, Iceland. Chem. Geol. 412, 34–47.

Michalczewska K., Hreinsdóttir S., Árnadóttir T., Hjartardóttir Á. R., Ágústsdóttir T.,

Gudmundsson M. T., Geirsson H., Sigmundsson F. and Gudmundsson G. B. (2012)

Inflation and deflation episdoes in the Krísuvík volcanic system. In American

Geophysical Union Fall Meeting

Nesbitt H. W. and Young G. M. (1982) Early Proterozoic climates and plate motion

inferred from major element chemistry of lutites. Nature 299, 715–717.

Park K.-H. (1990) Sr, Nd, and Pb isotope studies of ocean island basalts: constraints on

their origin and evolution. Columbia University.

Peate D. W., Baker J. A., Jakobsson S. P., Waight T. E., Kent A. J. R., Grassineau N. V.

and Skovgaard A. C. (2009) Historic magmatism on the Reykjanes Peninsula,

Iceland: a snap-shot of melt generation at a ridge segment. Contrib. to Mineral.

Petrol. 157, 359–382.

Peate D. W., Breddam K., Baker J. A., Kurz M. D., Barker A. K., Prestvik T., Grassineau

N. V. and Skovgaard A. C. (2010) Compositional Characteristics and Spatial

Distribution of Enriched Icelandic Mantle Components. J. Petrol. 51, 1447–1475.

Pedersen R. and Sigmundsson F. (2004) InSAR based sill model links spatially offset

areas of deformation and seismicity for the 1994 unrest episode at Eyjafjallajökull

volcano, Iceland. 31, 5–9.

Pedersen R. and Sigmundsson F. (2006) Temporal development of the 1999 intrusive

episode in the Eyjafjallajökull volcano, Iceland, derived from InSAR images. Bull.

Volcanol. 68, 377–393.

126

Raffone N., Ottolini L., Tonarini S., Gianelli G. and Fridleifsson G. Ó. (2008) A SIMS

study of lithium, boron and chlorine in basalts from Reykjanes (southwestern

Iceland). Microchim. Acta 161, 307–312.

Rudge J. F., Maclennan J. and Stracke A. (2013) The geochemical consequences of

mixing melts from a heterogeneous mantle. Geochim. Cosmochim. Acta 114, 112–

143.

Shorttle O., Maclennan J. and Piotrowski A. M. (2013) Geochemical provincialism in the

Iceland plume. Geochim. Cosmochim. Acta 122, 363–397.

Sigmarsson O. (1996) Short magma chamber residence time at an Icelandic volcano

inferred from U-series disequilibria. Nature 382, 440–442.

Sigmarsson O., Condomines M. and Fourcade S. (1992) A detailed Th, Sr and O isotope

study of Hekla: differentiation processes in an Icelandic volcano. Contrib. to

Mineral. Petrol. 112, 20–34.

Sigmarsson O., Condomines M., Grönvold K. and Thordarson T. (1991) Extreme magma

homogeneity in the 1783-1784 Lakagigar eruption: Origin of a large volume of

evolved basalt in Iceland. Geophys. Res. Lett. 18, 2229–2232.

Sigmarsson O., Vlastelic I., Andreasen R., Bindeman I., Devidal J.-L., Moune S.,

Keiding J. K., Larsen G., Höskuldsson A. and Thordarson T. (2011) Remobilization

of silicic intrusion by mafic magmas during the 2010 Eyjafjallajökull eruption. Solid

Earth 2, 271–281.

Sigmundsson F., Hreinsdóttir S., Hooper A., Árnadóttir T., Pedersen R., Roberts M. J.,

Óskarsson N., Auriac A., Decriem J., Einarsson P., Geirsson H., Hensch M.,

Ofeigsson B. G., Sturkell E., Sveinbjörnsson H. and Feigl K. L. (2010) Intrusion

triggering of the 2010 Eyjafjallajökull explosive eruption. Nature 468, 426–30.

Skovgaard A. C., Storey M., Baker J. A., Blusztajn J. and Hart S. R. (2001) Osmium-

oxygen isotopic evidence for a recycled and strongly depleted component in the

Iceland mantle plume. Earth Planet. Sci. Lett. 194, 259–275.

Stracke A., Zindler A., Salters V. J. M., McKenzie D., Blichert-Toft J., Albarède F. and

Grönvold K. (2003) Theistareykir revisited. Geochemistry, Geophys. Geosystems 4.

Sturkell E. and Sigmundsson F. (2003) Recent unrest and magma movements at

Eyjafjallajökull and Katla volcanoes, Iceland. J. Geophys. Res. 108.

Sun S. -s. and McDonough W. F. (1989) Chemical and isotopic systematics of oceanic

basalts: implications for mantle composition and processes. Geol. Soc. London,

Spec. Publ. 42, 313–345.

127

Taisne B., Tait S. and Jaupart C. (2011) Conditions for the arrest of a vertical propagating

dyke. Bull. Volcanol. 73, 191–204.

Tarasewicz J., Brandsdóttir B., White R. S. and Hensch M. (2012) Using

microearthquakes to track repeated magma intrusions beneath the Eyjafjallajökull

stratovolcano, Iceland. J. Geophys. Res. 117.

Tarasewicz J., White R. S., Brandsd B. and Schoonman C. M. (2014) Seismogenic

magma intrusion before the 2010 eruption of Eyjafjallajökull volcano, Iceland.

Geophys. J. Int. 198, 906–921.

Tarasewicz J., White R. S., Woods A. W., Brandsdóttir B. and Gudmundsson M. T.

(2012) Magma mobilization by downward-propagating decompression of the

Eyjafjallajökull volcanic plumbing system. Geophys. J. Int. 39.

Thirlwall M. F., Gee M. A. M., Taylor R. N. and Murton B. J. (2004) Mantle components

in Iceland and adjacent ridges investigated using double-spike Pb isotope ratios.

Geochim. Cosmochim. Acta 68, 361–386.

Todt W., Cliff R. A., Hanser A. and Hofmann A. W. (1993) Recalibration of NBS lead

standards using a 202Pb-205Pb double spike. Terra Abstr. 5, 396.

Appendix

Isotope geochemistry procedures

A.1 Acid digestion

Approximately 200 mg of each sample were weighed and placed in Teflon vials.

After adding 1 ml of HNO3 and 2 ml of HF, the samples were capped and left on a

hotplate for 5 days. Samples were dried down and 1 ml of HNO3 was added to each vial

and dried once more to remove any traces of HF. Between 2-4 ml of HCl were added to

the samples, capped to facilitate dissolution, and once again dried down on the hotplate.

Samples were dissolved in 1 ml of 0.7N HBr, then loaded into 2 ml centrifuge tubes. The

solutions were run for 3 minutes at a rate of 8400 rpm in the centrifuge.

A.1.1 Pb separation

To prepare the Pb columns, fresh DOWEX 200-400 mesh resin was added to each

columns and rinsed twice with 1 ml of 7N HCl to clean the resin. This was followed by

two 1 ml rinses of triple distilled water (referred to as “3 sub” as the third distillation is

under sub-boiling conditions). Another 1 ml of 7N HCl was added to the columns.

Subsequently, a final 1 ml of 3 sub water was added. After cleaning, the columns were

conditioned with 1 ml of 0.7N HBr was added to them. The acid collected from column

washing was disposed and the sample vials were loaded beneath the columns.

Samples were loaded carefully so as to avoid loading any of the residuum in the

centrifuge tubes. The columns were eluted with 1 ml of 0.7N HBr three times and once

129

more with 0.75 ml. Following a 0.3 ml elution of 3N HCl, the eluate was collected for Sr

and Nd and moved to a hot plate to dry down.

New sample vials were placed below the columns for Pb collection. Two 0.75 ml

elutions of 7N HCl were collected and these were placed on the hot plate to dry.

A.1.2 Sr and REE separation

The Sr columns with [NAME] resin were cleaned and conditioned with 2 sets of

2.5 ml of 2.5N HCl. Samples dissolved in 2.5N HCl were loaded in the columns and

subsequently rinsed three times with approximately 1/3 ml of 2.5N HCl. Either 12 or 16

ml of 2.5N HCl were added to the columns, depending on the amount of resin in the

column. The waste was disposed and Sr collection vials were placed below the columns.

Samples were eluted with 5 ml of 2.5N HCl and the eluate was collected Sr.

Either 2 or 3 ml of 6.6N HCl were added to each column based on the volume of

resin. The eluate was disposed. Rare earth element vials were placed below the columns.

Columns were eluted with 6 ml of 6.6N HCl for REEs and Nd separation. Both the Sr

and REE eluates were placed on the hot plate to dry.

A.1.3 Nd separation

Acid in the pre-conditioned columns was drained. Meanwhile, 1 ml of 0.26N HCl

was added to the dried REE sample. The bottom of the vial was gentle scratched and

stirred and with a pipette tip to facilitate dissolution. The samples were added to the

columns and subsequently rinsed with 1 ml of 0.26N HCl. Depending on the volume of

resin, 10-12 ml of 0.26N HCl was eluted and disposed. The Nd vials were placed under

the columns and 5 ml of 0.26N HCl were eluted and collected for Nd. Samples were

placed on the hot plate to dry.

130

A.2 Filament preparation

For Pb, a 99.999% Re ribbon was used for the filaments while 99.98% Re was

used for Sr and Nd. A single ribbon is welded onto a two-pronged sample holder for Pb

and Sr. For Nd analysis, two ribbons are welded onto a four-pronged sample holder. One

ribbon is used to host the sample while the other is oriented perpendicular and acts as

another ionization source behind the sample. Once welded to the sample holders, the Re

filaments were degassed and surface contamination was removed by running a 4.0 A

current through the filament.

A.2.1 Pb loading and analysis

Samples for Pb analysis were dissolved in 1 μl of H3PO4 and loaded onto the

filament between two layers of 1 μl tetraethyl orthosilicate (TEOS). During loading,

filaments had currents of 0.8-0.9 A. Once the filaments were dry, the current was

increased very briefly, 1-2 seconds, to 2.2 A. This procedure was also followed for the Pb

standards (NIST SRM 981). This standard was always analyzed at the beginning of a

batch of analyses and after every 3-4 samples. Samples were gradually heated to 1500

mA or ~900°C from 1000 mA at a rate of 100 mA/min. The current was thereafter

increased 30 mA/min towards a target of 1300°C. Ten blocks of 10 cycles were measured

for each sample once a stable signal was achieved.